Hostname: page-component-7857688df4-7g6pc Total loading time: 0 Render date: 2025-11-18T13:58:42.492Z Has data issue: false hasContentIssue false

SIZE OF PIECES IN DECOMPOSITIONS INTO THE FIRST UNCOUNTABLE CARDINAL MANY PIECES

Part of: Set theory

Published online by Cambridge University Press:  09 October 2025

WILLIAM CHAN*
Affiliation:
VIENNA UNIVERSITY OF TECHNOLOGY AUSTRIA
STEPHEN JACKSON
Affiliation:
UNIVERSITY OF NORTH TEXAS UNITED STATES E-mail: stephen.jackson@unt.edu, nam.trang@unt.edu
NAM TRANG
Affiliation:
UNIVERSITY OF NORTH TEXAS UNITED STATES E-mail: stephen.jackson@unt.edu, nam.trang@unt.edu
Rights & Permissions [Opens in a new window]

Abstract

Within the determinacy setting, ${\mathscr {P}({\omega _1})}$ is regular (in the sense of cofinality) with respect to many known cardinalities and thus there is substantial evidence to support the conjecture that ${\mathscr {P}({\omega _1})}$ has globally regular cardinality. However, there is no known information about the regularity of ${\mathscr {P}(\omega _2)}$. It is not known if ${\mathscr {P}(\omega _2)}$ is even $2$-regular under any determinacy assumptions. The article will provide the following evidence that ${\mathscr {P}(\omega _2)}$ may possibly be ${\omega _1}$-regular: Assume $\mathsf {AD}^+$. If $\langle A_\alpha : \alpha < {\omega _1} \rangle $ is such that ${\mathscr {P}(\omega _2)} = \bigcup _{\alpha < {\omega _1}} A_\alpha $, then there is an $\alpha < {\omega _1}$ so that $\neg (|A_\alpha | \leq |[\omega _2]^{<\omega _2}|)$.

Information

Type
Article
Creative Commons
Creative Common License - CCCreative Common License - BY
This is an Open Access article, distributed under the terms of the Creative Commons Attribution licence (https://creativecommons.org/licenses/by/4.0/), which permits unrestricted re-use, distribution, and reproduction in any medium, provided the original work is properly cited.
Copyright
© The Author(s), 2025. Published by Cambridge University Press on behalf of The Association for Symbolic Logic

1 Introduction

A cardinality is an equivalence class under the bijection relation on the class of a sets. The cardinality of X is denoted $|X|$ and consists of all sets in bijection with X. Cardinalities are ordered by the injection comparison relation: $|X| \leq |Y|$ if and only if there is an injection of X into Y. A cardinal is an ordinal which does not inject into any smaller ordinals. Assuming the axiom of choice, every cardinality has a unique cardinal as a member. The axiom of choice will not be assumed here.

If $\kappa $ is a cardinal, then the classical definition of the cofinality of $\kappa $ is ${\mathrm {cof}}(\kappa )$ is the least cardinal $\delta $ so that there is an increasing function $\rho : \delta \rightarrow \kappa $ so that $\sup (\rho ) = \kappa $ . An equivalent definition is that it is the least ordinal $\delta $ so that for all $\gamma < \delta $ and function $\Phi : \kappa \rightarrow \gamma $ , there is an $\alpha \in \gamma $ so that $|\Phi ^{-1}[\{\alpha \}]| = \kappa $ .

In choiceless settings, cardinalities no longer have unique cardinal members since sets may not well-orderable. The collection of cardinalities are also no longer well-ordered by the injection comparison relation. In [Reference Chan, Jackson and Trang8], the authors developed a robust notion of regularity and cofinality in the choiceless setting.

Let X be a set and Y be a class. X is said to have Y-regular cardinality if and only if for every function $\Phi : X \rightarrow Y$ , there is a $y \in Y$ so that $|\Phi ^{-1}[\{y\}]| = |X|$ . A set X is said to be locally regular if and only if for all sets Y with $|Y| < |X|$ , X has Y-regular cardinality. A set X is said to be globally regular if and only if for all sets Y which are surjective images of X and $\neg (|X| \leq |Y|)$ , X has Y-regular cardinality.

Since cardinalities are not well-ordered under the injection comparison relation, the natural definition of the cofinality of a set is formally a proper class:

  • The local cofinality of a set X is the class

    $$ \begin{align*}\mathsf{lcof}(X) &= \{Y : (\exists Z)(|Z| = |Y| \wedge Z \subseteq X \wedge X \text{ does not have } Y\text{-regular}\\ & \qquad \text{cardinality})\}.\end{align*} $$
  • Let $\mathsf {Surj}(X)$ be the class of all sets onto which X surjects. The global cofinality of a set X is the class

    $$ \begin{align*}\mathsf{gcof}(X) = \{Y \in \mathrm{Surj}(X) : X \text{ does not have } Y\text{-regular cardinality}\}.\end{align*} $$

Observe that if X has locally regular cardinality, then $\mathsf {lcof}(X) = |X|$ and if X has globally regular cardinality, then $\mathsf {gcof}(X) = \{Y \in \mathsf {Surj}(X) : |X| \leq |Y|\}$ .

The following summarizes some of the results obtained by the authors in [Reference Chan, Jackson and Trang8] concerning regularity and cofinality. If $\alpha $ is an ordinal, then $\mathsf {lcof}(\alpha ) = \{X : |{\mathrm {cof}}(\alpha )| \leq |X| \leq |\alpha |\}$ and $\mathsf {gcof}(\alpha ) = \{X \in \mathsf {Surj}(\alpha ) : |{\mathrm {cof}}(\alpha )| \leq |X|\}$ . Thus $\mathsf {lcof}(\alpha ) = \mathsf {gcof}(\alpha )$ . If $\kappa $ is a regular cardinal, then $\kappa $ has globally regular cardinality and $\mathsf {lcof}(\kappa ) = \mathsf {gcof}(\kappa ) = |\kappa |$ . Thus the choiceless theory of regularity and cofinality for well-orderable sets has a strong resemblance to the usual theory of cofinality in the choiceful framework.

Assuming $\mathsf {AC}^{\mathbb {R}}_{\omega }$ and all sets of reals have the perfect set property, $\mathbb {R}$ has locally regular cardinality and $\mathsf {lcof}(\mathbb {R}) = |\mathbb {R}|$ . Under $\mathsf {AD}^+$ , Woodin’s perfect dichotomy [Reference Caicedo and Ketchersid3, Reference Chan6] implies that $\mathbb {R}$ has globally regular cardinality and $\mathsf {gcof}(\mathbb {R}) = \{X \in \mathsf {Surj}(\mathbb {R}) : X \text { is not well-orderable}\}$ .

$E_0$ is the equivalence relation on ${{}^\omega 2}$ defined by $x \ E_0 \ y$ if and only if there exists an $m \in \omega $ so that for all $n \in \omega $ , if $m \leq n < \omega $ , then $x(n) = y(n)$ . Under $\mathsf {AD}^+$ , Hjorth’s dichotomy [Reference Hjorth12] implies that is globally regular and .

Under $\mathsf {AC}^{\mathbb {R}}_{\omega }$ and all subsets of $\mathbb {R}$ have the property of Baire and the perfect set property, $|\mathbb {R}|$ and $|{\omega _1}|$ are incomparable cardinalities. This can be used to show that $\mathbb {R} \sqcup {\omega _1}$ does not have $2$ -regular cardinalities. Thus $\mathsf {gcof}(\mathbb {R} \sqcup {\omega _1}) = \{X \in \mathsf {Surj}(\mathbb {R}) : |X| \geq 2\}$ . Under the same assumptions, $\mathbb {R} \times {\omega _1}$ does not have $\mathbb {R}$ -regular cardinality and does not have ${\omega _1}$ -regular cardinality. Under $\mathsf {AD}^+$ , the Woodin perfect set dichotomy will show that $\mathsf {gcof}(\mathbb {R} \times {\omega _1}) = \{X \in \mathsf {Surj}(\mathbb {R}) : X \text { is uncountable}\}$ .

Martin showed that ${\omega _1} \rightarrow _* ({\omega _1})^{\omega _1}_{<{\omega _1}}$ and $\omega _2 \rightarrow _* (\omega _2)^{<\omega _2}_{<\omega _2}$ under $\mathsf {AD}$ . The partition properties on ${\omega _1}$ can be used to show that for all $\epsilon \leq {\omega _1}$ , $[{\omega _1}]^\epsilon $ has $\omega $ -regular cardinality. If $\epsilon < {\omega _1}$ , then $[{\omega _1}]^\epsilon $ does not have ${\omega _1}$ -regular cardinality since ${[{\omega _1}]^\epsilon = \bigcup _{\delta < {\omega _1}} [\delta ]^{\omega _1}}$ by the regularity of ${\omega _1}$ and since $|[\delta ]^\epsilon | \leq |\mathbb {R}| < |[{\omega _1}]^\epsilon |$ . The partition relation on $\omega _2$ can be used to show that for all $\epsilon < \omega _2$ , $[\omega _2]^\epsilon $ has ${\omega _1}$ -regular cardinality. If $\epsilon < \omega _2$ , $[\omega _2]^\epsilon = \bigcup _{\delta < \omega _2} [\delta ]^\epsilon $ and hence as before, $[\omega _2]^\epsilon $ does not have $\omega _2$ -regular cardinality.

The strong partition property ${\omega _1} \rightarrow _* ({\omega _1})^{\omega _1}_2$ can be used to show that for each ${\lambda < {\omega _1}}$ , $[{\omega _1}]^{<{\omega _1}}$ has $\lambda $ -regular cardinality. $[{\omega _1}]^{<{\omega _1}}$ does not have ${\omega _1}$ -regular cardinality since $[{\omega _1}]^{<{\omega _1}} = \bigcup _{\epsilon < {\omega _1}} [{\omega _1}]^\epsilon $ and $|[{\omega _1}]^\epsilon | < |[{\omega _1}]^{<{\omega _1}}|$ for all $\epsilon < {\omega _1}$ .

At the present time, the regular cardinals, $\mathbb {R}$ , and are the only known locally or globally regular cardinalities. ${\mathscr {P}({\omega _1})}$ is the most natural candidate for another globally regular cardinality. The most important conjecture concerning regularity and cofinality is that ${\mathscr {P}({\omega _1})}$ has globally regular cardinality. [Reference Chan, Jackson and Trang8] has amassed substantial evidence that ${\mathscr {P}({\omega _1})}$ should be globally regular under determinacy assumptions. ${\mathscr {P}({\omega _1})}$ is regular with respect to essentially every set (which does not already have an injective copy of ${\mathscr {P}({\omega _1})}$ ) for which one currently has a practical understanding: [Reference Chan5] showed that ${\omega _1} \rightarrow _* ({\omega _1})^{\omega _1}_2$ implies that ${\mathscr {P}({\omega _1})}$ has $\mathrm {ON}$ -regular cardinality. One of the main results of [Reference Chan, Jackson and Trang8] is that ${\omega _1} \rightarrow _* ({\omega _1})^{{\omega _1}}_{<{\omega _1}}$ implies that ${\mathscr {P}({\omega _1})}$ has ${}^{<{\omega _1}}\mathrm {ON}$ -regular cardinality. (It is open if the strong partition property ${\omega _1} \rightarrow _* ({\omega _1})^{\omega _1}_2$ implies the very strong partition property ${\omega _1} \rightarrow _* ({\omega _1})^{\omega _1}_{<{\omega _1}}$ ; however, the very strong partition property on ${\omega _1}$ is a consequence of $\mathsf {AD}$ .) Assuming $\mathsf {AD}^+$ , ${\mathscr {P}({\omega _1})}$ is regular with respect to quotient of many familiar Borel equivalence relations. If E is an equivalence relation with all classes countable, then ${\mathscr {P}({\omega _1})}$ has -regular cardinality. If E is $E_0$ , $E_1$ , $E_2$ , a countable Borel equivalence relation, an essentially countable equivalence relation, a hyperfinite equivalence relation, a hypersmooth equivalence relation, or more generally a $\boldsymbol {\Sigma }^1_1$ equivalence relation which is pinned in any model of $\mathsf {ZFC}$ (in the sense of Zapletal [Reference Larson and Zapletal21]), then ${\mathscr {P}({\omega _1})}$ has -regular cardinality. The Friedman–Stanley jump of $=^+$ is not a pinned equivalence relation. Its quotient is in bijection with $\mathscr {P}_{\omega _1}(\mathbb {R})$ , the set of countable subsets of $\mathbb {R}$ . One can still show that ${\mathscr {P}({\omega _1})}$ has $\mathscr {P}_{\omega _1}(\mathbb {R})$ -regular cardinality under $\mathsf {AD}^+$ .

As mentioned above, $[\omega _2]^{<\omega _2}$ does not have $\omega _2$ -regular cardinality. Intuitively, one would expect $[\omega _2]^{<\omega _2}$ to at least have ${\omega _1}$ -regular cardinality. Above, it was remarked that the strong partition property ${\omega _1} \rightarrow _* ({\omega _1})^{\omega _1}_2$ implies $[{\omega _1}]^{<{\omega _1}}$ has $\omega $ -regular cardinality. However, $\omega _2$ is a weak but non-strong partition cardinal and thus the argument for $[{\omega _1}]^{<{\omega _1}}$ does not apply for $[\omega _2]^{<\omega _2}$ . Similarly, the intuition is that ${\mathscr {P}(\omega _2)}$ should be highly regular and perhaps globally regular.

However since $\omega _2$ is weak partition cardinal which is not a strong partition cardinal, $[\omega _2]^{<\omega _2}$ and ${\mathscr {P}(\omega _2)}$ seems just out of reach of the partition arguments and Martin’s ultrapower analysis of $\omega _2$ . (Surprisingly, $[\omega _2]^{<\omega _2}$ and more generally $[\omega _n]^{<\omega _2}$ for $2 \leq n < \omega $ can still be analyzed through the ultrapowers by measures on ${\omega _1}$ as shown in [Reference Chan, Jackson and Trang8]) Unlike ${\mathscr {P}({\omega _1})}$ , nothing is known about the cofinality of ${\mathscr {P}(\omega _2)}$ . For example, one does not know if ${\mathscr {P}(\omega _2)}$ even has $2$ -regular cardinality. The goal of this article is to produce some evidence that $[\omega _2]^{<\omega _2}$ and ${\mathscr {P}(\omega _2)}$ could have $2$ -regular cardinality or more generally could have ${\omega _1}$ -regular cardinality. (In the forthcoming [Reference Chan, Jackson and Trang8], the authors have shown that $[\omega _2]^{<\omega _2}$ and even $[\omega _n]^{<\omega _2}$ are ${\omega _1}$ -regular for all $2 \leq n < \omega $ .)

If $[\omega _2]^{<\omega _2}$ does not have ${\omega _1}$ -regular cardinality, then one can decompose $[\omega _2]^{<\omega _2}$ into an ${\omega _1}$ -length sequence of disjoint sets $\langle A_\alpha : \alpha < {\omega _1}\rangle $ so that $|A_\alpha | < |[\omega _2]^{<\omega _2}|$ . Although the structure of the cardinalities below $[\omega _2]^{<\omega _2}$ is far from understood, perhaps the largest natural cardinality of combinatorial flavor strictly below $[\omega _2]^{<\omega _2}$ is $[\omega _2]^{\omega _1}$ . An instance of ${\omega _1}$ -regularity for $[\omega _2]^{<\omega _2}$ would be to show that $[\omega _2]^{<\omega _2}$ cannot be a union of ${\omega _1}$ -many sets $\langle A_\alpha : \alpha < {\omega _1}\rangle $ so that each $|A_\alpha | \leq |[\omega _2]^{\omega _1}|$ .

Perhaps the largest natural cardinality strictly below ${\mathscr {P}(\omega _2)}$ is $|[\omega _2]^{<\omega _2}|$ . An instance of ${\omega _1}$ -regularity for ${\mathscr {P}(\omega _2)}$ would be to show that ${\mathscr {P}(\omega _2)}$ cannot be a union of ${\omega _1}$ -many sets $\langle A_\alpha : \alpha < {\omega _1}\rangle $ so that each $|A_\alpha | \leq |[\omega _2]^{<\omega _2}|$ .

The main results of this article will verify these two instances of $\omega _1$ -regularity:

  • (Theorem 3.18) Assume $\mathsf {AD}^+$ . If $\langle A_\alpha : \alpha < {\omega _1}\rangle $ is such that $[\omega _2]^{<\omega _2} = \bigcup _{\alpha < {\omega _1}} A_\alpha $ , then there exists an $\alpha < {\omega _1}$ so that $\neg (|A_\alpha | \leq |[\omega _2]^{\omega _1}|)$ .

  • (Theorem 3.19) Assume $\mathsf {AD}^+$ . If $\langle A_\alpha : \alpha < {\omega _1}\rangle $ is such that ${\mathscr {P}(\omega _2)} = \bigcup _{\alpha < {\omega _1}} A_\alpha $ , then there exists an $\alpha < {\omega _1}$ so that $\neg (|A_\alpha | \leq |[\omega _2]^{<\omega _2}|)$ .

Recently, the authors in [Reference Chan, Jackson and Trang8] have fully verified under $\mathsf {AD}$ the conjecture that $[\omega _2]^{<\omega _2}$ is ${\omega _1}$ -regular: For any $\langle A_\alpha : \alpha < {\omega _1}\rangle $ such that $[\omega _2]^{<\omega _2} = \bigcup _{\alpha < {\omega _1}} A_\alpha $ , there is an $\alpha < {\omega _1}$ so that $|A_\alpha | = |[\omega _2]^{<\omega _2}|$ . (More generally, for all $2 \leq n < \omega $ , $[\omega _n]^{<\omega _2}$ is ${\omega _1}$ -regular.) The verification of ${\omega _1}$ -regularity for $[\omega _2]^{<\omega _2}$ (or more generally, $[\omega _n]^{<\omega _2}$ when $2 \leq n < \omega $ ) uses a very technical analysis of the ultrapower of ${\omega _1}$ by the club ultrafilter on ${\omega _1}$ where the type or length of a function into $\omega _2$ represented by a function $f : {\omega _1} \rightarrow {\omega _1}$ is not fixed by varies with f. It is still not known if ${\mathscr {P}(\omega _2)}$ is $2$ -regular.

For each $1 \leq n < \omega $ , the projective ordinal $\boldsymbol {\delta }^1_n$ is the supremum of the length of $\boldsymbol {\Delta }^1_n$ prewellorderings on $\mathbb {R}$ . It can be shown that for all $n \in \omega $ , $\boldsymbol {\delta }^1_{2n + 2} = (\boldsymbol {\delta }^1_{2n + 1})^+$ . $\boldsymbol {\delta }^1_1 = {\omega _1}$ and $\boldsymbol {\delta }^1_2 = \omega _2$ . Also $\boldsymbol {\delta }^1_3 = \omega _{\omega + 1}$ and $\boldsymbol {\delta }^1_{4} = \omega _{\omega + 2}$ . The last section will show that the results for ${\omega _1}$ and $\omega _2$ can be generalized to each odd projective ordinal $\boldsymbol {\delta }^1_{2n + 1}$ and the next even projective ordinal $\boldsymbol {\delta }^1_{2n + 2}$ .

  • (Theorem 4.39) Assume $\mathsf {AD}^+$ . Let $n \in \omega $ . If $\langle A_\alpha : \alpha < \boldsymbol {\delta }^1_{2n + 1}\rangle $ is such that ${\mathscr {P}(\boldsymbol {\delta }^1_{2n + 2})} = \bigcup _{\alpha < \boldsymbol {\delta }^1_{2n + 1}} A_\alpha $ , then there is an $\alpha < \boldsymbol {\delta }^1_{2n + 1}$ so that $\neg (|A_\alpha | \leq |[\boldsymbol {\delta }^1_{2n + 2}]^{<\boldsymbol {\delta }^1_{2n + 2}}|)$ .

2 Cardinality of sets of functions on ordinals

$\mathsf {ZF}$ will be assumed throughout and all additional assumptions will be made explicit.

Definition 2.1. If X and Y are sets, then let ${}^X Y$ be the set of all functions from X to Y.

If $\delta $ is an ordinal and X is a set, then let ${}^{<\delta }X = \bigcup _{\epsilon < \delta } {}^\epsilon X$ .

If $\delta $ and $\lambda $ are ordinals and $X \subseteq \lambda $ , then let $[X]^\delta $ be the collection of all increasing functions $f : \delta \rightarrow X$ . Let $[X]^{<\delta } = \bigcup _{\epsilon < \delta } [X]^{\epsilon }$ .

If $\delta $ is a cardinal and X is a set, then let $\mathscr {P}_{\delta }(X) = \{A \in {\mathscr {P}(X)} : |A| < \delta \}$ .

If $\delta \leq \lambda $ are ordinals, then let $IB(\delta ,\lambda ) = \{f \in {}^\delta \lambda : (\forall \alpha < \delta )(\sup (f \upharpoonright \alpha ) < \lambda )\}$ .

This section collects some basic results concerning the cardinality of sets of the form $[\lambda ]^\delta $ , ${}^\delta \lambda $ , and $[\lambda ]^{<\delta }$ .

Fact 2.2. Let $\delta \leq \lambda $ be ordinals such that $\delta $ is a cardinal. Then $|[\lambda ]^{<\delta }| = |\mathscr {P}_\delta (\lambda )| = |{}^{<\delta }\lambda |$ .

Proof. Let $\Phi : [\lambda ]^{<\delta } \rightarrow \mathscr {P}_\delta (\lambda )$ be defined by $\Phi (f) = \mathrm {rang}(f)$ . $\Phi $ is a bijection.

Let $\pi : \lambda \times \lambda \rightarrow \lambda $ be a bijection. For $f \in {}^{<\delta }\lambda $ , let $G_f = \{\pi (\alpha ,\beta ) : \alpha \in \mathrm {dom}(f) \wedge f(\alpha ) = \beta \}$ . Note that since $\mathrm {dom}(f) \in \delta $ and $\delta $ is a cardinal, $|G_f| < \delta $ . Thus $G_f \in \mathscr {P}_{\delta }(\lambda )$ . Define $\Psi : {}^{<\delta }\lambda \rightarrow \mathscr {P}_\delta (\lambda )$ by $\Psi (f) = G_f$ . $\Psi $ is an injection. The previous paragraph showed that there is a bijection of $\mathscr {P}_\delta (\lambda )$ into $[\lambda ]^{<\delta }$ and $[\lambda ]^{<\delta } \subseteq {}^{<\delta }\lambda $ . Thus there is an injection $\Psi : \mathscr {P}_\delta (\lambda ) \rightarrow {}^{<\delta }\lambda $ . By the Cantor–Schröder–Bernstein theorem, $|{}^{<\delta }\lambda | = |\mathscr {P}_\delta (\lambda )| = |[\lambda ]^{<\delta }|$ .

Say an ordinal $\lambda $ is indecomposable if and only if for all $\alpha , \beta < \lambda $ , $\alpha + \beta < \lambda ,$ and $\alpha \cdot \beta < \lambda $ .

Fact 2.3. If $\delta \leq \lambda $ are ordinals and $\lambda $ is indecomposable, then $|IB(\delta ,\lambda )| = |[\lambda ]^\delta |$ .

Proof. For $f \in IB(\delta ,\lambda )$ , define $\Phi (f) \in [\lambda ]^\delta $ by recursion as follows. Suppose for all $\beta < \delta $ , $\Phi (f) \upharpoonright \beta $ has been defined and for all $\alpha < \beta $ , $\Phi (f)(\alpha ) \leq \sup (f \upharpoonright \alpha + 1) \cdot (\alpha + 1) < \lambda $ . Then $\sup (\Phi (f) \upharpoonright \beta ) \leq \sup (f \upharpoonright \beta ) \cdot \beta < \lambda $ since $\sup (f \upharpoonright \beta ) < \lambda $ and $\lambda $ is indecomposable. Let $\Phi (f)(\beta ) = \sup (\Phi (f) \upharpoonright \beta ) + f(\beta ),$ which is less than $\lambda $ since $\lambda $ is indecomposable. Then $\Phi (f)(\beta ) = \sup (\Phi (f) \upharpoonright \beta ) + f(\beta ) \leq \sup (f \upharpoonright \beta ) \cdot \beta + f(\beta ) \leq \sup (f \upharpoonright \beta + 1) \cdot (\beta + 1) < \lambda $ since $\lambda $ is indecomposable.

This defines $\Phi : IB(\delta ,\lambda ) \rightarrow [\lambda ]^\delta $ . Note that for all $\alpha < \delta $ , $f(\alpha )$ is the unique ordinal $\gamma $ so that $\Phi (f)(\alpha ) = \sup (\Phi (f) \upharpoonright \alpha ) + \gamma $ . Thus $\Phi $ is an injection. Thus $|IB(\delta ,\lambda )| \leq |[\lambda ]^\delta |$ . Since $[\lambda ]^\delta \subseteq IB(\delta ,\lambda )$ , $|[\lambda ]^\delta | \leq |IB(\delta ,\delta )|$ . By the Cantor–Schröder–Bernstein, $|[\lambda ]^\delta | = |IB(\delta ,\lambda )|$ .

Fact 2.4. Let $\delta \leq \lambda $ be ordinals such that $\lambda $ is indecomposable and $\delta \leq {\mathrm {cof}}(\lambda )$ . Then $|{}^\delta \lambda | = |[\lambda ]^\delta |$ .

Proof. Suppose $\delta \leq {\mathrm {cof}}(\lambda )$ . For all $f \in {}^\delta \lambda $ and $\alpha < \delta $ , $\sup (f \upharpoonright \alpha ) < \lambda $ . Thus ${}^\delta \lambda \subseteq B(\delta ,\lambda )$ . Thus $|{}^\delta \lambda | = |B(\delta ,\lambda )| = |[\lambda ]^\delta |$ by Fact 2.3.

Fact 2.5. Let $\delta \leq \lambda $ be ordinals such that $\lambda $ is indecomposable, ${\mathrm {cof}}(\delta ) = {\mathrm {cof}}(\lambda )$ , and $\delta < {\mathrm {cof}}(\lambda )^+$ . Then $|{}^\delta \lambda | = |[\lambda ]^\delta |$ .

Proof. Note that $|{}^\delta \lambda | = |{}^{{\mathrm {cof}}(\lambda )}\lambda |$ since $|\delta | = |{\mathrm {cof}}(\delta )|$ . By Fact 2.4, $|{}^{{\mathrm {cof}}(\lambda )}\lambda | = |[\lambda ]^{{\mathrm {cof}}(\lambda )}|$ . Thus $|{}^\delta \lambda | = |[\lambda ]^{{\mathrm {cof}}(\lambda )}|$ . Thus it suffices to produce an injection of $[\lambda ]^{{\mathrm {cof}}(\lambda )}$ into $[\lambda ]^\delta $ . Let $\rho : {\mathrm {cof}}(\lambda ) \rightarrow \delta $ . Since $\lambda $ is indecomposable, $\delta \cdot \lambda = \lambda $ . For each $\alpha < \lambda $ , let $\iota (\alpha )$ be the least $\beta < {\mathrm {cof}}(\lambda )$ so that $\alpha \leq \rho (\beta )$ . For $f \in [\lambda ]^{{\mathrm {cof}}(\lambda )}$ , let $\Phi (f) : \delta \rightarrow \lambda $ be defined by $\Phi (f)(\alpha ) = \delta \cdot f(\iota (\alpha )) + \alpha $ . One can check that for all $f \in [\lambda ]^{{\mathrm {cof}}(\lambda )}$ , $\Phi (f) \in [\lambda ]^\delta $ and $\Phi : [\lambda ]^{{\mathrm {cof}}(\lambda )} \rightarrow [\lambda ]^\delta $ is an injection.

Fact 2.6. If $\kappa $ is a measurable cardinal (has a $\kappa $ -complete nonprincipal ultrafilter on $\kappa $ ), then for all $\delta < \kappa $ , there is no injection of $\kappa $ into ${\mathscr {P}(\delta )}$ .

Proof. Suppose $\Phi : \kappa \rightarrow {\mathscr {P}(\delta )}$ is a function. Let $\mu $ be a $\kappa $ -complete nonprincipal ultrafilter on $\kappa $ . For each $\alpha < \delta $ and $i \in \{0,1\}$ , let $A^i_\alpha = \{\beta < \kappa : \Phi (\beta )(\alpha ) = i\}$ (where elements of ${\mathscr {P}(\delta )}$ are identified with elements of ${}^\delta 2$ ). For each $\alpha < \delta $ , let $i_\alpha $ be the unique $i \in \{0,1\}$ so that $A_\alpha ^{i_\alpha } \in \mu $ . Since $\mu $ is $\kappa $ -complete, $\bigcap _{\alpha < \delta } A_\alpha ^{i_\alpha } \in \mu $ . Let $f \in {}^\delta 2$ be defined by $f(\alpha ) = i_\alpha $ . Since $\mu $ is nonprincipal, let $\alpha _1 < \alpha _2 < \delta $ so that $\alpha _1,\alpha _2 \in \bigcap _{\alpha < \delta } A_\alpha ^{i_\alpha }$ . $\Phi (\alpha _1) = f = \Phi (\alpha _2)$ . Thus $\Phi $ is not an injection.

Under $\mathsf {AD}$ , ${\omega _1}$ is a strong partition cardinal and $\omega _2$ is a weak partition cardinal. Thus ${\omega _1}$ and $\omega _2$ are measurable cardinals. More generally, $\boldsymbol {\delta }^1_{2n + 1}$ is a strong partition cardinal and $\boldsymbol {\delta }^1_{2n + 2}$ is a weak partition cardinal. (It is known that $\boldsymbol {\delta }^1_{3} = \omega _{\omega + 1}$ and $\boldsymbol {\delta }^1_4 = \omega _{\omega + 2}$ .) (See [Reference Chan6, Reference Ketchersid18] or [Reference Kleinberg19] for more information concerning partition properties under $\mathsf {AD}$ and the associated measures.)

If $\kappa $ is a cardinal, then one says boldface $\mathsf {GCH}$ holds at $\kappa $ if and only if there is no injection of $\kappa ^+$ into ${\mathscr {P}(\kappa )}$ . Boldface $\mathsf {GCH}$ holds below $\kappa $ if and only if boldface $\mathsf {GCH}$ holds at all $\delta < \kappa $ . Fact 2.6 implies the following result.

Fact 2.7. Assume $\mathsf {AD}$ . Boldface $\mathsf {GCH}$ holds at $\omega $ and $\omega _1$ .

Recently, the authors [Reference Chan, Jackson and Trang7] have shown boldface $\mathsf {GCH}$ below $\omega _\omega $ using purely combinatorial arguments assuming $\mathsf {AD}$ (in fact, just assuming ${\omega _1} \rightarrow ({\omega _1})^{\omega _1}_2$ and the ultrapower of ${\omega _1}$ by the club measure on ${\omega _1}$ is $\omega _2$ ). Steel [Reference Steel25, Reference Steel and Woodin26] showed that if $L(\mathbb {R}) \models \mathsf {AD}$ , then $L(\mathbb {R})\models $ “boldface $\mathsf {GCH}$ holds below $\Theta $ ” using an inner model theoretic analysis of $\mathrm {HOD}$ . Thus by the Moschovakis coding lemma, it is a theorem of $\mathsf {AD}$ that boldface $\mathsf {GCH}$ holds below $\Theta ^{L(\mathbb {R})}$ . More generally, Woodin showed that $\mathsf {AD}^+$ implies the boldface $\mathsf {GCH}$ holds below $\Theta $ .

Fact 2.8. Suppose $\lambda $ is cardinal and $\lambda $ does not inject into ${\mathscr {P}(\kappa )}$ for any $\kappa < \lambda $ . Then $\neg (|[\lambda ]^{{\mathrm {cof}}(\lambda )}| \leq |\bigcup _{\delta \leq \kappa < \lambda } [\kappa ]^\delta |)$ .

Proof. Suppose there is an injection $\Phi : [\lambda ]^{{\mathrm {cof}}(\lambda )} \rightarrow \bigcup _{\delta \leq \kappa < \lambda }[\kappa ]^\delta $ . Let $\tilde \Phi \subseteq [\lambda ]^{{\mathrm {cof}}(\lambda )} \times \lambda \times \lambda $ be defined by $(f,\alpha ,\beta ) \in \tilde \Phi $ if and only if $\alpha \in \mathrm {dom}(\Phi (f))$ and $\Phi (f)(\alpha ) = \beta $ . $L[\tilde \Phi ] \models \mathsf {ZFC}$ . In $L[\tilde \Phi ]$ , define $\Psi : [\lambda ]^{{\mathrm {cof}}(\lambda )} \rightarrow \bigcup _{\delta \leq \kappa < \lambda } [\kappa ]^\delta $ by $\Psi (f)(\alpha ) = \beta $ if and only if $\tilde \Phi (f,\alpha ,\beta )$ . Note $\Psi \in L[\tilde \Phi ]$ and $L[\tilde \Phi ] \models \Psi : [\lambda ]^{{\mathrm {cof}}(\lambda )} \rightarrow \bigcup _{\delta \leq \kappa < \lambda } [\kappa ]^\delta $ is an injection. If there are $\delta \leq \kappa < \lambda $ so that $L[\tilde \Phi ] \models \lambda \leq |[\kappa ]^\delta |$ , then there is an injection of $\lambda $ into $[\kappa ]^\delta \subseteq {\mathscr {P}(\kappa )}$ in the real world. This contradicts the assumption that $\lambda $ does not inject into ${\mathscr {P}(\kappa )}$ for any $\kappa < \lambda $ . Thus $L[\tilde \Phi ] \models |\bigcup _{\delta \leq \kappa < \lambda } [\kappa ]^\delta | = \lambda $ . By a theorem of $\mathsf {ZFC}$ , $L[\tilde \Phi ] \models |[\lambda ]^{{\mathrm {cof}}(\lambda )}| \geq \lambda ^+$ . It is impossible that $L[\tilde \Phi ] \models \Psi : [\lambda ]^{{\mathrm {cof}}(\lambda )} \rightarrow \bigcup _{\delta \leq \kappa < \lambda } [\kappa ]^\delta $ is an injection.

Fact 2.9. Suppose $\kappa $ is a regular cardinal and there is no injection of $\kappa $ into ${\mathscr {P}(\delta )}$ for any $\delta < \kappa $ . Then $|[\kappa ]^{<\kappa }| < |{\mathscr {P}(\kappa )}|$ .

Proof. It is clear that $|[\kappa ]^{<\kappa }| \leq |{\mathscr {P}(\kappa )}|$ . Since $\kappa $ is regular, $[\kappa ]^{<\kappa } = \bigcup _{\delta \leq \mu < \kappa } [\mu ]^\delta $ . By Fact 2.8, $\neg (|{\mathscr {P}(\kappa )}| = |[\kappa ]^\kappa | \leq |\bigcup _{\delta \leq \mu < \kappa } [\mu ]^\delta ]| = [\kappa ]^{<\kappa })$ .

Since Martin showed that $\omega _2 \rightarrow (\omega _2)^2_2$ (and in fact, $\omega _2 \rightarrow (\omega _2)^{\epsilon }_2$ for all $\epsilon < \omega _2$ ), $\omega _2$ is a regular cardinal.

Fact 2.10. Assume $\mathsf {AD}$ . $|[\omega _2]^{<\omega _2}| < |{\mathscr {P}(\omega _2)}|$ .

Proof. This follows from Facts 2.7 and 2.9.

Fact 2.11. Let $\delta \leq \lambda $ be ordinals such that ${\mathrm {cof}}(\lambda ) < {\mathrm {cof}}(\delta )$ and $\lambda $ does not inject into ${\mathscr {P}(\kappa )}$ for all $\kappa < \lambda $ . Then $|[\lambda ]^\delta | < |{}^\delta \lambda |$ .

Proof. It is clear that $[\lambda ]^\delta \subseteq {}^\delta \lambda $ . Since ${\mathrm {cof}}(\delta ) \neq {\mathrm {cof}}(\lambda )$ , $[\lambda ]^{\delta } = \bigcup _{\kappa < \lambda }[\kappa ]^\delta \subseteq \bigcup _{\mu \leq \kappa < \lambda }[\kappa ]^\mu $ . Define $\Psi : [\lambda ]^{{\mathrm {cof}}(\lambda )} \rightarrow {}^\delta \lambda $ by

$$ \begin{align*}\Psi(f)(\alpha) = \begin{cases} f(\alpha) & \quad \alpha < {\mathrm{cof}}(\lambda) \\ 0 & \quad {\mathrm{cof}}(\lambda) < \alpha \end{cases}.\end{align*} $$

$\Psi $ is an injection. Thus if there was an injection of ${}^{\delta }\lambda $ into $|[\lambda ]^\delta |$ , then there would be an injection of $[\lambda ]^{{\mathrm {cof}}(\lambda )}$ into $\bigcup _{\mu \leq \kappa < \lambda } [\kappa ]^\mu $ , which contradicts Fact 2.8.

Example 2.12. Assume $\mathsf {AD}$ . Recall Steel showed the boldface $\mathsf {GCH}$ holds below $\Theta ^{L(\mathbb {R})}$ (and one can directly use the analysis of the ultrapower by the finite partition measures on ${\omega _1}$ to show the boldface $\mathsf {GCH}$ below $\omega _{\omega + 1}$ ).

  1. (1) $|[\omega _\omega ]^{{\omega _1}}| < |{}^{\omega _1}\omega _\omega |$ . This follows from Fact 2.11. The cardinality of the collection of the increasing sequences can be smaller than the cardinality of the collection of all sequences.

  2. (2) $|IB({\omega _1},\omega _\omega )| = |[\omega _\omega + \omega ]^{{\omega _1}}| < |IB({\omega _1}, \omega _\omega + \omega )| = |{}^{\omega _1}(\omega _\omega + \omega )|$ . To see this: Note that $[\omega _\omega + \omega ]^{\omega _1} = [\omega _\omega ]^{\omega _1} \subseteq \bigcup _{\delta \leq \kappa < \omega _\omega } [\kappa ]^\delta $ . Thus by Fact 2.8, $[\omega _\omega ]^{\omega _1}$ does not inject into $\bigcup _{\delta \leq \kappa < \omega _\omega } [\kappa ]^\delta $ and thus does not inject into $[\omega _\omega + \omega ]^{\omega _1}$ . However $[\omega _\omega ]^{\omega _1} \subseteq IB({\omega _1},\omega _\omega + \omega )$ . This shows that $|[\omega _\omega + \omega ]^{\omega _1}| < |IB({\omega _1},\omega _\omega + \omega )|$ . Notice that $\omega _\omega + \omega $ is not indecomposable. This shows that the indecomposability assumption of Fact 2.3 is necessary. Also since $[\omega _\omega + \omega ]^{{\omega _1}} = [\omega _\omega ]^{\omega _1}$ , $|[\omega _\omega + \omega ]^{\omega _1}| = |[\omega _\omega ]^{\omega _1}| = |IB({\omega _1},\omega _\omega )|$ by Fact 2.3. Note that ${}^{\omega _1}(\omega _\omega ) \subseteq IB({\omega _1},\omega _\omega + \omega ) \subseteq {}^{\omega _1}(\omega _{\omega } + \omega )$ and $|{}^{\omega _1}(\omega _\omega + \omega )| = |{}^{\omega _1} \omega _\omega |$ . Thus $|{}^{\omega _1}(\omega _\omega + \omega )| = |IB({\omega _1},\omega _{\omega } + \omega )|$ . This shows that $|[\omega _\omega + \omega ]^{\omega _1}| < |{}^{\omega _1}(\omega _\omega + \omega )|$ .

Fact 2.13.

  • [Reference Chan, Jackson and Trang9] $(\mathsf {AD}) [{\omega _1}]^{<{\omega _1}}$ does not inject into ${}^\omega (\omega _\omega )$ .

  • [Reference Chan, Jackson and Trang9] $(\mathsf {AD} + \mathsf {DC}_{\mathbb {R}})$ $[{\omega _1}]^{<{\omega _1}}$ does not inject into ${}^\omega \mathrm {ON}$ , the class of $\omega $ -sequences of ordinals.

  • [Reference Chan, Jackson and Trang10] More generally, if $\kappa \rightarrow (\kappa )^{<\kappa }_2$ ( $\kappa $ is a weak partition cardinal), then $[\kappa ]^{<\kappa }$ does not inject into ${}^\lambda \mathrm {ON}$ , for all $\lambda < \kappa $ .

Fact 2.14. Assume $\mathsf {AD}$ . $|[\omega _2]^{\omega _1}| < |[\omega _2]^{<\omega _2}|$ .

Proof. Under $\mathsf {AD}$ , Martin showed that $\omega _2$ is a weak partition cardinal (that is, satisfies $\omega _2 \rightarrow _* (\omega _2)^{<\omega _2}_2$ ). The result follows from the third point in Fact 2.13.

Example 2.15. Assume $\mathsf {AD}$ . Note that $\neg (|[\omega _\omega ]^\omega | \leq |[\omega _\omega ]^{\omega _1}|)$ .This is because if there was an injection of $[\omega _\omega ]^\omega $ into $[\omega _\omega ]^{\omega _1}$ , then there would be an injection of $[\omega _\omega ]^\omega $ into $[\omega _\omega ]^{\omega _1} = \bigcup _{{\omega _1} \leq \kappa < \omega _\omega } [\kappa ]^{\omega _1} \subseteq \bigcup _{\delta \leq \kappa < \omega _\omega } [\kappa ]^\delta $ (where the first equality follows from the fact that $[\omega _\omega ]^{\omega _1}$ consists of increasing ${\omega _1}$ -sequences and ${\mathrm {cof}}(\omega _\omega ) = \omega $ ), which violates Fact 2.8. Note that $\neg (|[\omega _\omega ]^{\omega _1}| \leq |[\omega _\omega ]^\omega |)$ . This is because $[{\omega _1}]^{<{\omega _1}}$ injects into $[\omega _\omega ]^{\omega _1}$ and $[{\omega _1}]^{<{\omega _1}}$ does not inject into ${}^\omega \mathrm {ON}$ by Fact 2.13. Since $[\omega _\omega ]^{\omega _1}$ injects into $[\omega _\omega ]^{{\omega _1} + \omega }$ , this shows that $|[\omega _\omega ]^\omega | < |[\omega _\omega ]^{{\omega _1} + \omega }|$ .

See [Reference Chan4] for more information concerning distinguishing sets of the form $[\kappa ]^\delta $ and ${}^\delta \kappa $ for varying $\delta \leq \kappa < \Theta $ under $\mathsf {AD}^+$ .

3 Decomposition into ${\omega _1}$ many pieces

Definition 3.1. Fix a bijection $\pi : \omega \times \omega \rightarrow \omega $ . If $x \in {{}^\omega \omega }$ and $k \in \omega $ , then let $x^{[k]} \in {{}^\omega \omega }$ be defined by $x^{[k]}(n) = x(\pi (k,n))$ .

If $x \in {{}^\omega 2}$ , then define $\mathcal {R}_x \subseteq \omega \times \omega $ by $\mathcal {R}_x(m,n)$ if and only if $x(\pi (m,n)) = 1$ . Let $\mathrm {field}(x) = \mathrm {field}(\mathcal {R}_x) = \{m : (\exists n)(\mathcal {R}_x(m,n) \vee \mathcal {R}_x(n,m))\}$ .

Let ${\mathrm {WO}} = \{w \in {{}^\omega 2} : \mathcal {R}_w \text { is a wellordering}\}$ . Let ${\mathrm {ot}} : {\mathrm {WO}} \rightarrow {\omega _1}$ be defined by ${\mathrm {ot}}(w)$ is the order type of $(\mathrm {field}(w), \mathcal {R}_w)$ . If $\alpha < {\omega _1}$ , then let ${\mathrm {WO}}_\alpha = \{w \in {\mathrm {WO}} : {\mathrm {ot}}(w) = \alpha \}$ .

Definition 3.2. Let $\alpha < {\omega _1}$ . For $s \in {}^{<\omega }\alpha $ , let $N^\alpha _s = \{f \in {}^\omega \alpha : s \subseteq f\}$ . Give ${}^\omega \alpha $ the topology generated by $\{N^\alpha _s : s \in {}^{<\omega }\alpha \}$ as a basis (which is the product of the discrete topology on $\alpha $ ). Then ${}^\omega \alpha $ is homeomorphic to ${{}^\omega \omega }$ with its usual topology.

Under $\mathsf {AD}$ , all subsets of ${{}^\omega \omega }$ have the Baire property and thus well ordered unions of meager subsets of ${{}^\omega \omega }$ are meager in ${{}^\omega \omega }$ . (For the latter fact: Given a well-ordered sequence of meager sets whose union is nonmeager, consider the horizontal and vertical section of the prewellordering induced by the sequence to obtain a contradiction.) Therefore under $\mathsf {AD}$ , for all $\alpha < {\omega _1}$ , all subsets of ${}^\omega \alpha $ have the Baire property and well-ordered unions of meager subsets of ${}^\omega \alpha $ are meager in ${}^\omega \alpha $ .

For $\alpha < {\omega _1}$ , let $\mathsf {surj}_\alpha = \{f \in {}^\omega {\omega _1} : f[\omega ] = \alpha \}$ . For all $\alpha < {\omega _1}$ , $\mathrm {surj}_\alpha $ is comeager in ${}^\omega \alpha $ .

If $\alpha < {\omega _1}$ , $p \in {}^{<\omega }\alpha $ , and $\varphi $ is a formula, then let $(\forall ^{*,\alpha }_p f)\varphi (f)$ be the assertion that for comeagerly many $f \in N_p^\alpha $ , $\varphi (f)$ holds.

Definition 3.3. For each $f \in {}^\omega {\omega _1}$ , let $A_f = \{n \in \omega : (\forall m < n)(f(m) \neq f(n))\}$ . (Note for all $f \in {}^\omega {\omega _1}$ , $f \upharpoonright A_f : A_f \rightarrow f[\omega ]$ is a bijection.)

For $f \in {}^\omega {\omega _1}$ , let $\mathfrak {G}(f) \in {{}^\omega 2}$ be defined by $\mathfrak {G}(f)(\pi (m,n)) = 1$ if and only if $m \in A_f$ , $n \in A_f$ , and $f(m) < f(n)$ . $\mathfrak {G}$ is a simple form of the Kechris–Woodin generic coding function for ${\omega _1}$ , which is developed more generally in [Reference Kechris and Woodin17].

Fact 3.4. $\mathfrak {G} : {}^\omega {\omega _1} \rightarrow {\mathrm {WO}}$ and for all $\alpha < {\omega _1}$ , if $f \in \mathsf {surj}_\alpha $ , then $\mathfrak {G}(f) \in {\mathrm {WO}}_\alpha $ .

Proof. Note that $(\mathrm {field}(\mathfrak {G}(f)), \mathcal {R}_{\mathfrak {G}(f)}) = (A_f, \mathcal {R}_{\mathfrak {G}(f)})$ is order isomorphic to $(f[A_f],<),$ where $<$ is the usual ordering on ${\omega _1}$ . Thus $\mathfrak {G}(f)$ does indeed belong to ${\mathrm {WO}}$ . Also if $f \in \mathsf {surj}_\alpha $ , then $f[A_f] = \alpha $ and thus $\mathfrak {G}(f) \in {\mathrm {WO}}_\alpha $ .

Definition 3.5. Let $\langle \rho _r : r \in \mathbb {R}\rangle $ be some standard coding of strategies ${\rho : {}^{<\omega }\omega \rightarrow \omega }$ on $\omega $ by reals. Let $\Xi _r : \mathbb {R} \rightarrow \mathbb {R}$ be the Lipschitz continuous function corresponding to the strategy $\rho _r$ . (That is, for each $f \in {}^\omega \omega $ , $\Xi _r(f) \in {}^\omega \omega $ is defined by recursion by $\Xi _r(f)(n) = \rho _r(\langle f(0), \Xi _r(f)(0), \ldots , f(n - 1),\Xi _r(f)(n - 1), f(n)\rangle )$ .) Note that $\langle \Xi _r : r \in \mathbb {R}\rangle $ is a coding of all Lipschitz continuous function by reals.

If $A, B \in \mathbb {R}$ , then write $A \leq _L B$ if and only if there is an $r \in \mathbb {R}$ so that ${A = \Xi _r^{-1}[B]}$ . The Wadge lemma under $\mathsf {AD}$ asserts that for all $A,B \in {\mathscr {P}(\mathbb {R})}$ , $A \leq _L B$ or ${(\mathbb {R}\setminus B) \leq _L A}$ .

Martin–Monk showed that under $\mathsf {AD}$ and $\mathsf {DC}_{\mathbb {R}}$ , $\leq _L$ is a wellfounded relation. For each $A \in {\mathscr {P}(\mathbb {R})}$ , let $\mathrm {rk}_L(A) \in \mathrm {ON}$ be the rank of A in $\leq _L$ . Let $\Theta $ be the supremum of the ordinals, which are surjective images of $\mathbb {R}$ . It can be shown that $\Theta $ is the length of $\leq _L$ and thus for all $A \in {\mathscr {P}(\mathbb {R})}$ , $\mathrm {rk}_L(A) < \Theta $ .

Fact 3.6. (Moschovakis coding lemma) Assume $\mathsf {AD}$ . Let $\Gamma $ be a pointclass closed under $\exists ^{\mathbb {R}}$ , $\wedge $ , and continuous preimages. Let $(P,\preceq )$ be a prewellordering in $\Gamma $ . Let $\kappa $ be the length of $(P,\preceq )$ and $\varphi : P \rightarrow \kappa $ be the associated surjective norm. If $R \subseteq P \times \mathbb {R}$ , then there is an $S \in \Gamma $ with the following property:

  • $S \subseteq R.$

  • For all $\alpha < \kappa $ , there exists a $p \in P$ and $x \in \mathbb {R}$ so that $\varphi (p) = \alpha $ and $R(p,x)$ if and only if there exists a $p \in P$ and $x \in \mathbb {R}$ so that $\varphi (p) = \alpha $ and $S(p,x)$ .

The following is a useful coarse consequence of the Moschovakis coding lemma.

Fact 3.7. If $\kappa $ is a surjective image of $\mathbb {R}$ (i.e., $\kappa < \Theta $ ), then $\mathbb {R}$ surjects onto ${\mathscr {P}(\kappa )}$ .

Fix the following notation which will be used in the discussion that follows: Let X be a surjective image of $\mathbb {R}$ . Fix a surjection $\pi : \mathbb {R} \rightarrow X$ . Let $\delta \leq \lambda < \Theta $ . By Fact 3.7, there is a surjection $\varpi : \mathbb {R} \rightarrow {\mathscr {P}(\lambda )}$ . If $B \subseteq \mathbb {R}$ , let $T_B = \{(x,f) : (\exists z \in B)(x = \pi (z^{[0]}) \wedge f = \varpi (z^{[1]})\}$ . Let $\langle A_\alpha : \alpha < \nu \rangle $ be such that for all $\alpha < \nu $ , $A_\alpha \subseteq X$ . (In this section, $\nu $ will either be $\omega $ or ${\omega _1}$ .) In the below applications, $|A_\alpha | \leq |{}^{<\delta }\lambda |$ or $|A_\alpha | \leq |{}^\delta \lambda |$ for all $\alpha < \nu $ . Elements of ${}^{<\delta }\lambda $ or ${}^\delta \lambda $ can be identified as elements of ${\mathscr {P}(\lambda \times \lambda )}$ or of ${\mathscr {P}(\lambda )}$ (after coding pairs). As an example, if $A \subseteq X$ and $\Phi : A \rightarrow {}^{<\delta }\lambda $ , then the graph of $\Phi $ is $T_B$ , where $B = \{z \in \mathbb {R} : \Phi (\pi (z^{[0]})) = \varpi (z^{[1]})\}$ .

Theorem 3.8. Assume $\mathsf {AD}$ . Suppose X is a surjective image of $\mathbb {R}$ . Let $\delta \leq \lambda $ be cardinals so that $1 \leq \delta < \Theta $ and $\omega \leq \lambda < \Theta $ . Let $\langle A_n : n \in \omega \rangle $ be a sequence so that for all $n \in \omega $ , $A_n \subseteq X$ . Assume one of the following three settings:

  1. (1) $|A_n | \leq | {}^{<\delta }\lambda |$ for all $n \in \omega $ .

  2. (2) $|A_n | \leq |{}^\delta \lambda |$ for all $n \in \omega $ .

  3. (3) $|A_n | \leq |[\lambda ]^\delta |$ for all $n \in \omega $ .

Assume that there is a $Z \in {\mathscr {P}(\mathbb {R})}$ so that for all $n \in \omega $ , there exists an $r \in \mathbb {R}$ so that $T_{\Xi _r^{-1}[Z]}$ is a graph of an injection of $A_n$ into ${}^{<\delta }\lambda $ in (1) (into ${}^\delta \lambda $ in (2) or $[\lambda ]^\delta $ in (3)). Then, respectively, the following hold:

  1. (1) $|\bigcup _{n \in \omega } A_n| \leq |{}^{<\delta }\lambda |$ .

  2. (2) $|\bigcup _{n \in \omega } A_n| \leq |{}^\delta \lambda |$ .

  3. (3) $|\bigcup _{n \in \omega } A_n| \leq |[\lambda ]^{\delta }|$ .

Proof. Assume the setting of (1) that for all $n \in \omega $ , $|A_n| \leq |{}^{<\delta }\lambda |$ . Let $R \subseteq \omega \times \mathbb {R}$ be defined by $R(n,r)$ if and only if $T_{\Xi ^{-1}_r[Z]}$ is the graph of an injection of $A_n$ into ${}^{<\delta }\lambda $ . (Recall that $\Xi ^{-1}_r[Z]$ is the subset of $\mathbb {R}$ Lipchitz reducible to Z via the Lipschitz continuous function $\Xi _r$ and $T_{\Xi ^{-1}_r[Z]}$ was defined before the statement of Theorem 3.8.) By $\mathsf {AC}^{\mathbb {R}}_\omega $ , there is a sequence $\langle r_n : n \in \omega \rangle $ so that for all $n \in \omega $ , $R(n, r_n)$ . Thus for all $n \in \omega $ , $T_{\Xi _{r_n}^{-1}[Z]}$ is the graph of an injection $A_n$ into ${}^{<\delta }\lambda $ . Let $\Phi _n : A_n \rightarrow {}^{<\delta }\lambda $ be the injection whose graph is $T_{\Xi _{r_n}^{-1}[Z]}$ . For each $x \in \bigcup _{n \in \omega } A_n$ , let $\iota (x)$ be the least n so that $x \in A_n$ . Since $\omega \leq \lambda $ , let $\varsigma : \omega \times \lambda \rightarrow \lambda $ be a bijection. Define $\Phi : \bigcup _{n \in \omega } A_n \rightarrow {}^{<\delta }\lambda $ by letting $\Phi (x) \in [\lambda ]^{|\Phi _{\iota (x)}(x)|}$ be defined by $\Phi (x)(\gamma ) = \varsigma (\iota (x),\Phi _{\iota (x)}(x)(\gamma ))$ . Suppose $x \neq y$ . If $\iota (x) \neq \iota (y)$ , then $\Phi (x) \neq \Phi (y)$ since $\varsigma $ is a bijection. If $\iota (x) = \iota (y)$ with common value $n \in \omega $ , then $\Phi _n(x) \neq \Phi _n(y)$ since $\Phi _n$ is an injection. Then again $\Phi (x) \neq \Phi (y)$ since $\varsigma $ is an injection. This establishes that $\Phi $ is an injection.

In the setting of (2), in which for all $n \in \omega $ , $|A_n| \leq |{}^\delta \lambda |$ , the proof is essentially the same.

In the setting of (3), in which for all $n \in \omega $ , $|A_n| \leq |[\lambda ]^\delta |$ , observe that the bijection $\varsigma : \omega \times \lambda \rightarrow \lambda $ may be chosen with the property that for all $n \in \omega $ and $\alpha < \beta < \lambda $ , $\varsigma (n,\alpha ) < \varsigma (n,\beta )$ . (For instance, $\varsigma $ derived from the Gödel pairing function would have such property.) Then the resulting function $\Phi (x)$ defined as above would belong to $[\lambda ]^\delta $ .

Theorem 3.9. Assume $\mathsf {AD}$ . Suppose X is a surjective image of $\mathbb {R}$ . Let $\langle A_\alpha : \alpha < {\omega _1} \rangle $ be a sequence so that for all $\alpha < {\omega _1}$ , $A_\alpha \subseteq X$ . Let $\delta $ and $\lambda $ be cardinals such that ${\omega _1} \leq \delta \leq \lambda < \Theta $ . Assume one of the following three settings:

  1. (1) ${\mathrm {cof}}(\delta ) \geq {\omega _1}$ and for all $\alpha < {\omega _1}$ , $|A_\alpha | \leq |{}^{<\delta }\lambda |$ .

  2. (2) For all $\alpha < {\omega _1}$ , $|A_\alpha | \leq |{}^\delta \lambda |$ .

  3. (3) ${\mathrm {cof}}(\lambda ) \geq {\omega _1}$ , and for all $\alpha < {\omega _1}$ , $|A_\alpha | \leq |[\lambda ]^\delta |$ .

Assume that there is a $Z \in {\mathscr {P}(\mathbb {R})}$ so that for all $\alpha < {\omega _1}$ , there exists an $r \in \mathbb {R}$ so that $T_{\Xi _r^{-1}[Z]}$ is the graph of an injection of $A_\alpha $ into $[\lambda ]^{<\delta }$ in (1) (into ${}^\delta \lambda $ in (2) or into $[\lambda ]^\delta $ in (3)). Then, respectively, the following hold:

  1. (1) $|\bigcup _{\alpha < {\omega _1}} A_\alpha | \leq |{}^{<\delta }\lambda |$ .

  2. (2) $|\bigcup _{\alpha < {\omega _1}} A_\alpha | \leq |{}^\delta \lambda |$ .

  3. (3) $|\bigcup _{\alpha < {\omega _1}} A_\alpha | \leq |[\lambda ]^\delta |$ .

Proof. Assume the setting of $(1)$ that for all $\alpha < {\omega _1}$ , $|A_\alpha | \leq |{}^{<\delta }\lambda |,$ where ${{\mathrm {cof}}(\delta ) \geq {\omega _1}}$ . Since $|{}^{<\delta } \lambda \setminus \{\emptyset \}| = |{}^{<\delta }\lambda |$ , injections from $A_\alpha $ into ${}^{<\delta }\lambda \setminus \{\emptyset \}$ will be considered to simplify notation.

Let ${\mathrm {WO}} \subseteq \mathbb {R}$ be the $\Pi _1^1$ set of reals coding wellorderings and ${\mathrm {ot}} : {\mathrm {WO}} \rightarrow {\omega _1}$ be the associated surjective norm given by the order type function. Define ${R \subseteq {\mathrm {WO}} \times \mathbb {R}}$ by $R(w,r)$ if and only if $T_{\Xi _r^{-1}[Z]}$ is the graph of an injection of $A_{{\mathrm {ot}}(w)}$ into ${{}^{<\delta }\lambda \setminus \{\emptyset \}}$ . $({\mathrm {WO}},{\mathrm {ot}})$ is a prewellordering which belongs to the pointclass $\boldsymbol {\Sigma }_2^1$ which is closed under continuous preimage, $\wedge $ , and $\exists ^{\mathbb {R}}$ . By the Moschovakis coding lemma (Fact 3.6), there is a $\boldsymbol {\Sigma }_2^1$ set $S \subseteq R$ so that for all $\alpha < {\omega _1}$ , there is a $w \in {\mathrm {WO}}_\alpha $ and $r \in \mathbb {R}$ so that $S(w,r)$ . Let $\leq _{\Pi _1^1} \in \Pi _1^1$ and $\leq _{\Sigma _1^1} \in \Sigma _1^1$ be the two norm relations which witness that $({\mathrm {WO}},{\mathrm {ot}})$ is a $\Pi _1^1$ -norm. Let $\tilde S(w,r)$ if and only if $w \in {\mathrm {WO}} \wedge (\exists v)(v \leq _{\Sigma _1^1} w \wedge w \leq _{\Sigma _1^1} v \wedge S(v,r))$ . $\tilde S \in \boldsymbol {\Sigma }_2^1$ and $\mathrm {dom}(\tilde S) = {\mathrm {WO}}$ . Since $\boldsymbol {\Sigma }_2^1$ has the scale property, let $\Lambda : {\mathrm {WO}} \rightarrow \mathbb {R}$ be a uniformization with the property that for all $w \in {\mathrm {WO}}$ , $\tilde S(w,\Lambda (w))$ . Thus for all $w \in {\mathrm {WO}}$ , $R(w,\Lambda (w))$ . For all $w \in {\mathrm {WO}}$ , $T_{\Xi ^{-1}_{\Lambda (w)}[Z]}$ is the graph of an injection of $A_{{\mathrm {ot}}(w)}$ into ${}^{<\delta }\lambda \setminus \{\emptyset \}$ . For each $w \in {\mathrm {WO}}$ , let $\Phi _w : A_{{\mathrm {ot}}(w)} \rightarrow {}^{<\delta }\lambda \setminus \{\emptyset \}$ be the injection whose graph is $T_{\Xi _{\Lambda (w)}^{-1}[Z]}$ .

For each $x \in \bigcup _{\alpha < {\omega _1}} A_\alpha $ , let $\iota (x)$ be the least $\alpha < {\omega _1}$ so that $x \in A_\alpha $ . Note that $|{}^{<\omega }{\omega _1}| = |{\omega _1}|$ . Let $\sigma : {\omega _1} \times {}^{<\omega }{\omega _1} \times \delta \times \lambda \rightarrow \lambda $ be a bijection. Define

$$ \begin{align*}\Upsilon(x) &= \{\sigma(\iota(x),p,\eta,\zeta) : (\exists \epsilon < \delta)(\forall^{*,\iota(x)}_p f)(\epsilon = \mathrm{dom}(\Phi_{\mathfrak{G}(f)}(x))\\& \qquad \wedge \eta < \epsilon \wedge \Phi_{\mathfrak{G}(f)}(x)(\eta) = \zeta)\}.\end{align*} $$

Observe that $\Upsilon (x) \in \mathscr {P}(\lambda )$ .

Fix $x \in \bigcup _{\alpha < {\omega _1}} A_\alpha $ . Let $K_x = \{p \in {}^{<\omega }\iota (x) : (\exists \eta ,\zeta )(\sigma (\iota (x),p,\eta ,\zeta ) \in \Upsilon (x)\}$ . If $p \in K_x$ , then there is a unique $\epsilon <\delta $ so that $(\forall ^{*,\iota (x)}_p f)(\mathrm {dom}(\Phi _{\mathfrak {G}(f)}(x)) = \epsilon )$ . To see this, suppose $\epsilon , \hat \epsilon < \delta $ are such that $(\forall ^{*,\iota (x)}_p f)(\mathrm {dom}(\Phi _{\mathfrak {G}(f)}(x)) = \epsilon )$ and $(\forall ^{*,\iota (x)}_p f)(\mathrm {dom}(\Phi _{\mathfrak {G}(f)}(x)) = \hat \epsilon )$ . Let $A_0 = \{f \in N_p^{\iota (x)} : \mathrm {dom}(\Phi _{\mathfrak {G}(f)}(x)) = \epsilon \}$ and $A_1 = \{f \in N_p^{\iota (x)} : \mathrm {dom}(\Phi _{\mathfrak {G}(f)}(x)) = \hat \epsilon \}$ . $A_0$ and $A_1$ are comeager subsets of $N_p^{\iota (x)}$ . Thus $A_0 \cap A_1 \neq \emptyset $ . Let $h \in A_0 \cap A_1$ . Then $\epsilon = \mathrm {dom}(\Phi _{\mathfrak {G}(h)}(x)) = \hat {\epsilon }$ . Let $\epsilon ^x_p$ be this unique $\epsilon $ associated with x and p. Let $U_{x,p} = \{\eta < \epsilon ^x_p : (\exists \zeta )(\sigma (\iota (x),p,\eta ,\zeta ) \in \Upsilon (x)\}$ . Note that $|U_{x,p}| \leq |\epsilon _p^x|$ . If $\eta \in U_{x,p}$ , there is a unique $\zeta $ such that $\sigma (\iota (x),p,\eta ,\zeta ) \in \Upsilon (x)$ . To see this, suppose $\zeta _1,\zeta _2$ so that $\sigma (\iota (x),p,\eta ,\zeta _1),\sigma (\iota (x),p,\eta ,\zeta _2) \in \Upsilon (x)$ . Then $B_0 = \{f \in N^{\iota (x)}_p : \Phi _{\mathfrak {G}(f)}(x)(\eta ) = \zeta _1\}$ and $B_1 = \{f \in N^{\iota (x)}_p : \Phi _{\mathfrak {G}(f)}(x)(\eta ) = \zeta _2\}$ are comeager in $N^{\iota (x)}_p$ . $B_0 \cap B_1$ is comeager in $N^{\iota (x)}_p$ . Let $h \in B_0 \cap B_1$ . Then $\zeta _1 = \Phi _{\mathfrak {G}(h)}(x)(\eta ) = \zeta _2$ . Let $\zeta ^x_{p,\eta }$ be this unique $\zeta $ . Thus $\Upsilon (x) = \{\sigma (\iota (x),p,\eta , \zeta ^x_{p,\eta }) : p \in K_x \wedge \eta \in U_{x,p}\}$ . Thus $|\Upsilon (x)| \leq |\bigcup _{p \in K_x} U_{x,p}| \leq \sup \{|\epsilon ^x_{p}| : p \in K_x\} < \delta $ since $|K_x| \leq |{}^{<\omega }\iota (x)| = \omega $ because $\iota (x) < {\omega _1}$ and ${\mathrm {cof}}(\delta )> \omega $ . Thus $\Upsilon (x)$ has cardinality less than $\delta $ and hence $\Upsilon (x) \in \mathscr {P}_\delta (\lambda )$ . It has been shown that $\Upsilon : \bigcup _{\alpha < {\omega _1}} A_\alpha \rightarrow \mathscr {P}_\delta (\lambda )$ .

Next, one will show that for all $x \in \bigcup _{\alpha < {\omega _1}} A_\alpha $ , $\Upsilon (x) \neq \emptyset $ . Let $\alpha = \iota (x)$ . Let $E_1 : \mathsf {surj}_\alpha \rightarrow \delta $ be defined by $E_1(f) = \mathrm {dom}(\Phi _{\mathfrak {G}(f)}(x))$ . Since well-ordered unions of meager subsets of ${}^\omega \alpha $ is a meager subset of ${}^\omega \alpha $ and $\mathsf {surj}_\alpha $ is a comeager subset of ${}^\omega \alpha $ , there is some $\epsilon < \delta $ so that $E_1^{-1}[\{\epsilon \}]$ is nonmeager. Let $E_2 : E_1^{-1}[\{\epsilon \}] \rightarrow \lambda $ be defined by $E_2(f) = \Phi _{\mathfrak {G}(f)}(x)(0)$ . Again since $E_1^{-1}[\{\epsilon \}]$ is nonmeager and well-ordered unions of meager sets are meager, there is some $\zeta < \lambda $ so that $E_2^{-1}[\{\zeta \}]$ is nonmeager. By the Baire property, there is a $p \in {}^{<\omega }\alpha $ so that $E_2^{-1}[\{\zeta \}]$ is comeager in $N_p^\alpha $ . Then $\sigma (\alpha ,p,0,\zeta ) \in \Upsilon (x)$ . $\Upsilon (x) \neq \emptyset $ .

Next, it will be shown that $\Upsilon $ is an injection. Suppose $x \neq y$ . First, suppose $\iota (x) \neq \iota (y)$ . Above, it was shown that $\Upsilon (x) \neq \emptyset $ . Let $\sigma (\iota (x),p,\eta ,\zeta ) \in \Upsilon (x)$ . Since $\sigma $ is an injection and all elements of $\Upsilon (y)$ take the form $\sigma (\iota (y),\hat {p},\hat {\eta },\hat {\zeta })$ , $\Upsilon (x) \neq \Upsilon (y)$ . Next, suppose that $\iota (x) = \iota (y)$ and denote this common ordinal by $\alpha $ . Let $D = \{f \in \mathsf {surj}_\alpha : \mathrm {dom}(\Phi _{\mathfrak {G}(f)}(x)) \neq \mathrm {dom}(\Phi _{\mathfrak {G}(f)}(y))\}$ . First suppose D is nonmeager. Consider $\varpi : D \rightarrow \delta \times \delta $ by $\varpi (f) = (\mathrm {dom}(\Phi _{\mathfrak {G}(f)}(x)), \mathrm {dom}(\Phi _{\mathfrak {G}(f)}(y)))$ . Since a well-ordered union of meager sets is meager and D is not meager, there is some $\epsilon _1,\epsilon _2 < \delta $ so that $\varpi ^{-1}[\{(\epsilon _1,\epsilon _2)\}]$ is nonmeager. Without loss of generality, suppose $\epsilon _1 < \epsilon _2$ . Define $\varsigma : \varpi ^{-1}[\{(\epsilon _1,\epsilon _2)\}] \rightarrow \lambda $ by $\varsigma (f) = \Phi _{\mathfrak {G}(f)}(y)(\epsilon _1)$ . Since $\varpi ^{-1}[\{(\epsilon _1,\epsilon _2)\}]$ is nonmeager and well-ordered union of meager sets is meager, there is a $\zeta \in \lambda $ so that $\varsigma ^{-1}[\{\zeta \}]$ is nonmeager. By the Baire property, let $p \in {}^{<\omega }\alpha $ be such that $\varsigma ^{-1}[\{\zeta \}]$ is comeager in $N^{\alpha }_p$ . Then $\sigma (\alpha ,p,\epsilon _1,\zeta ) \in \Upsilon (y)$ . However, $\sigma (\alpha ,p,\epsilon _1,\zeta ) \notin \Upsilon (x)$ since $(\forall ^{*,\alpha }_p f)(\mathrm {dom}(\Phi _{\mathfrak {G}(f)}(x)) = \epsilon _1)$ . In this case, $\Upsilon (x) \neq \Upsilon (y)$ . Finally, suppose ${}^\omega \alpha \setminus D$ is comeager. Let $\Sigma : {}^\omega \alpha \setminus D \rightarrow \delta $ be defined by $\Sigma (f) = \mathrm {dom}(\Phi _{\mathfrak {G}(f)}(x)) = \mathrm {dom}(\Phi _{\mathfrak {G}(f)}(y))$ . Since ${}^\omega \alpha \setminus D$ is comeager, there is some $\epsilon < \delta $ so that $\Sigma ^{-1}[\{\epsilon \}]$ is nonmeager. Note that since $\Phi _{\mathfrak {G}(f)}$ is an injection for all $f \in \mathsf {surj}(\alpha )$ (because it is the injection whose graph is $T_{\Xi ^{-1}_{\Lambda (\mathfrak {G}(f))}}[Z]$ ), $\Phi _{\mathfrak {G}(f)}(x) \neq \Phi _{\mathfrak {G}(f)}(y)$ . Define $\Pi : \Sigma ^{-1}[\{\epsilon \}] \rightarrow \epsilon $ be defined by $\Pi (f)$ is the least $\eta < \epsilon $ so that $\Phi _{\mathfrak {G}(f)}(x)(\eta ) \neq \Phi _{\mathfrak {G}(f)}(y)(\eta )$ . Since $\Sigma ^{-1}[\{\epsilon \}]$ is nonmeager, there is an $\eta < \epsilon $ so that $\Pi ^{-1}[\{\eta \}]$ is nonmeager. Let $\Gamma : \Pi ^{-1}[\{\eta \}] \rightarrow \lambda \times \lambda $ be defined by $\Gamma (f) = (\Phi _{\mathfrak {G}(f)}(x)(\eta ), \Phi _{\mathfrak {G}(f)}(y)(\eta ))$ . Since $\Pi ^{-1}[\{\eta \}]$ is nonmeager, there are $\zeta _1, \zeta _2 \in \lambda $ with $\zeta _1 \neq \zeta _2$ so that $\Gamma ^{-1}[\{(\zeta _1,\zeta _2)\}]$ is nonmeager. Since all subsets of ${}^\omega \alpha $ have the Baire property, there is a $p \in {}^{<\omega }\alpha $ so that $\Gamma ^{-1}[\{(\zeta _1,\zeta _2)\}] $ is comeager in $N^{\alpha }_p$ . Then $\sigma (\alpha ,p,\eta ,\zeta _1) \in \Upsilon (x)$ and $\sigma (\alpha ,p,\eta ,\zeta _1) \notin \Upsilon (y)$ . Thus $\Upsilon (x) \neq \Upsilon (y)$ . It has been shown that $\Upsilon : \bigcup _{\alpha < {\omega _1}} A_\alpha \rightarrow \mathscr {P}_\delta (\lambda )$ is an injection. Fact 2.2 shows $|{}^{<\delta }\lambda | = |\mathscr {P}_\delta (\lambda )|$ .

Next assume the setting of (2). The following will sketch the necessary modifications. By the same argument as above, for each $w \in {\mathrm {WO}}$ , there is an injection $\Phi _w : A_{{\mathrm {ot}}(w)} \rightarrow {}^\delta \lambda $ . Let

$$ \begin{align*}K_x = \{(p,\eta) : p \in {}^{<\omega}\iota(x) \wedge \eta < \delta \wedge (\exists \zeta < \lambda)(\forall^{*,\iota(x)}_p f)(\Phi_{\mathfrak{G}(f)}(x)(\eta) = \zeta)\}.\end{align*} $$

For each $(p,\eta ) \in K_x$ , by the argument provided above, there is a unique $\zeta $ so that $(\forall ^{*,\iota (x)}_p f)(\Phi _{\mathfrak {G}(f)}(x)(\eta ) = \zeta )$ . Thus for each $(p,\eta ) \in K_x$ , let $\zeta ^x_{p,\eta }$ be this unique $\zeta $ . Note that $K_x \subseteq {}^{<\omega }\iota (x) \times \delta \subseteq {}^{<\omega } {\omega _1} \times \delta $ . Let $\tau : {}^{<\omega }{\omega _1} \times \delta \rightarrow \delta $ be a bijection. Let $\mu : {\omega _1} \times \lambda \rightarrow \lambda $ be a bijection. Define $\Upsilon : X \rightarrow {}^\delta \lambda $ by

$$ \begin{align*}\Upsilon(x)(\alpha) = \begin{cases} \mu(\iota(x),0) & \quad \tau^{-1}(\alpha) \notin K_x \\ \mu(\iota(x),\zeta^x_{p,\eta}) & \quad \tau^{-1}(\alpha) \in K_x \wedge \tau^{-1}(\alpha) = (p,\eta) \end{cases}.\end{align*} $$

Finally, one will to show $\Upsilon $ is an injection. Suppose $x,y \in \bigcup _{\alpha < {\omega _1}} A_\alpha $ and $x \neq y$ . If $\iota (x) \neq \iota (y)$ , then $\Upsilon (x) \neq \Upsilon (y)$ since $\mu $ is a bijection. Now suppose $\iota (x) = \iota (y)$ and let $\alpha $ denote this common ordinal. For all $f \in \mathsf {surj}_\alpha $ , $\Phi _{\mathfrak {G}(f)}(x) \neq \Phi _{\mathfrak {G}(f)}(y)$ . Let $\Sigma : \mathsf {surj}_\alpha \rightarrow \delta $ be defined by $\Sigma (f)$ is the least $\eta < \delta $ so that $\Phi _{\mathfrak {G}(f)}(x)(\eta ) \neq \Phi _{\mathfrak {G}(f)}(y)(\eta )$ . Since $\mathsf {surj}_\alpha $ is comeager in ${}^\omega \alpha $ and well-ordered unions of meager sets are meager, there is an $\eta < \delta $ so that $\Sigma ^{-1}[\{\eta \}]$ is nonmeager. Let $\Pi : \Sigma ^{-1}[\{\eta \}] \rightarrow \lambda \times \lambda $ be defined by $\Pi (f) = (\Phi _{\mathfrak {G}(f)}(x)(\eta ), \Phi _{\mathfrak {G}(f)}(y)(\eta ))$ . Since $\Sigma ^{-1}[\{\eta \}]$ is nonmeager, there is some $\zeta _1,\zeta _2 < \lambda $ so that $\zeta _1 \neq \zeta _2$ and $\Pi ^{-1}[\{(\zeta _1,\zeta _2)\}]$ is nonmeager. By the Baire property, let $p \in {}^{<\omega }\alpha $ so that $\Pi ^{-1}[\{(\zeta _1,\zeta _2)\}]$ is comeager in $N^\alpha _p$ . Let $\beta = \tau (p,\eta )$ . Then $\Upsilon (x)(\beta ) = \mu (\alpha , \zeta _1) \neq \mu (\alpha , \zeta _2) = \Upsilon (y)(\beta )$ . Thus $\Upsilon (x) \neq \Upsilon (y)$ . It has been shown that $\Upsilon $ is an injection.

Assume the setting of (3). Let $K_x$ , $\zeta ^x_{p,\eta }$ , and $\tau : {}^{<\omega }{\omega _1} \times \delta \rightarrow \delta $ be defined as in (2). The bijection $\mu : {\omega _1} \times \lambda \rightarrow \lambda $ can be chosen with the property that for all $\nu < {\omega _1}$ and $\gamma < \lambda $ , $\sup \{\mu (\nu ,\beta ) : \beta < \gamma \} < \lambda $ . Let $\Upsilon $ be defined as above in (2). For $x \in X$ , $\gamma < \delta $ , and $p \in {}^{<\omega }\iota (x)$ , let $P^x_{\gamma ,p} = \{\eta \in \delta : \tau (p,\eta ) < \gamma \wedge \tau (p,\eta ) \in K_x\}$ . For each $p \in {}^{<\omega }\iota (x)$ , let $F^x_{p,\gamma } = \{\zeta _{p,\eta }^x : \eta \in P^x_{\gamma ,p}\}$ . The claim is that $F^x_{p,\gamma }$ is bounded below $\lambda $ . To see this, suppose $F^x_{p,\gamma }$ is not bounded below $\lambda $ . For each $\eta \in P^x_{\gamma ,p}$ , let $Y^x_{p,\gamma ,\eta } = \{f \in N^{\iota (x)}_p : \Phi _{\mathfrak {G}(f)}(x)(\eta ) = \zeta ^x_{p,\eta }\}$ . Each $Y^x_{p,\gamma ,\eta }$ is comeager in $N^{\iota (x)}_p$ . Since well-ordered intersection of comeager subsets of $N^{\iota (x)}_p$ is comeager in $N^{\iota (x)}_p$ , $\bigcap _{\eta \in P^x_{\gamma , p}} Y^x_{p,\gamma ,\eta }$ is comeager in $N^{\iota (x)}_p$ and is in particular nonempty. Let $f \in \bigcap _{\eta \in P^x_{\gamma , p}} Y^x_{p,\gamma ,\eta }$ . Then $\sup (\Phi _{\mathfrak {G}(f)}(x) \upharpoonright \gamma ) \geq \sup \{\zeta ^x_{p,\eta } : \eta \in P^x_{\gamma ,p}\} = \sup (F^x_{p,\gamma }) = \lambda $ . Then since $\gamma < \delta $ , $\Phi _{\mathfrak {G}(f)}(x)(\gamma ) \geq \lambda $ and hence $\Phi _{\mathfrak {G}(f)}(x) \notin [\lambda ]^\delta $ . This is a contradiction. Thus for all $p \in {}^{<\omega }\iota (x)$ , $\sup (F^x_{p,\gamma }) < \lambda $ . Since ${\mathrm {cof}}(\lambda )> \omega $ and $|{}^{<\omega }\iota (x)| = \omega $ , $\sup \{\sup (F^x_{p,\gamma }) : p \in {}^{<\omega }\iota (x)\} < \lambda $ . Note that $\sup (\Upsilon (x) \upharpoonright \gamma ) \leq \sup \{\mu (\iota (x),\zeta ) : \zeta \in \bigcup _{p \in {}^{<\omega }\iota (x)} F^x_{p,\gamma }\} \leq \sup \{\mu (\iota (x),\zeta ) : \zeta < \sup \{\sup (F^x_{p,\gamma }) : p \in {}^{<\omega }\iota (x)\}\} < \lambda $ (by the property of chosen bijection $\mu $ ). This shows that $\Upsilon : \bigcup _{\alpha < {\omega _1}} A_\alpha \rightarrow IB(\delta ,\lambda )$ . $\Upsilon $ is an injection by the same argument as in (2). The result now follows from Fact 2.3.

Theorem 3.10. Assume $\mathsf {AD}$ , $\mathsf {DC}_{\mathbb {R}}$ , and ${\mathrm {cof}}(\Theta )> {\omega _1}$ . Let X be a surjective image of $\mathbb {R}$ . Let $\langle A_\alpha : \alpha < {\omega _1}\rangle $ be a sequence so that for all $\alpha < {\omega _1}$ , $A_\alpha \subseteq X$ . Let $\delta $ and $\lambda $ be cardinals so that ${\omega _1} \leq \delta \leq \lambda < \Theta $ . Assume one of the following three settings:

  1. (1) ${\mathrm {cof}}(\delta ) \geq {\omega _1}$ and for all $\alpha < {\omega _1}$ , $|A_\alpha | \leq |{}^{<\delta }\lambda |$ .

  2. (2) For all $\alpha < {\omega _1}$ , $|A_\alpha | \leq |{}^\delta \lambda |$ .

  3. (3) ${\mathrm {cof}}(\lambda ) \geq {\omega _1}$ and for all $\alpha < {\omega _1}$ , $|A_\alpha | \leq |[\lambda ]^\delta |$ .

Then, respectively, the following hold:

  1. (1) $|\bigcup _{\alpha < {\omega _1}} A_\alpha | \leq |{}^{<\delta }\lambda |$ .

  2. (2) $|\bigcup _{\alpha < {\omega _1}} A_\alpha | \leq |{}^\delta \lambda |$ .

  3. (3) $|\bigcup _{\alpha < {\omega _1}} A_\alpha | \leq |[\lambda ]^\delta |$ .

Proof. Recall that Martin and Monk showed that $\mathsf {AD}$ and $\mathsf {DC}_{\mathbb {R}}$ prove that strict Lipschitz reduction is wellfounded. For each $\alpha < {\omega _1}$ , let $\beta _\alpha $ be the least $\beta $ so that there is some $B \in {\mathscr {P}(\mathbb {R})}$ with $\mathrm {rk}_L(B) = \beta $ and $T_{B}$ is the graph of an injection of $A_\alpha $ into ${}^{<\delta }\lambda $ . Since ${\mathrm {cof}}(\Theta )> {\omega _1}$ , $\sup \{\beta _\alpha : \alpha < {\omega _1}\} < \Theta $ . Let $Z \in {\mathscr {P}(\mathbb {R})}$ so that $\mathrm {rk}_L(Z) = \sup \{\beta _\alpha : \alpha < {\omega _1}\}$ . The result now follows from Theorem 3.9.

Theorem 3.11. Assume $\mathsf {AD}$ , $\mathsf {DC}_{\mathbb {R}}$ , and ${\mathrm {cof}}(\Theta )> \omega $ . Suppose X is a surjective image of $\mathbb {R}$ . Let $1 \leq \delta < \Theta $ and $\omega \leq \lambda < \Theta $ . Let $\langle A_n : n \in \omega \rangle $ be a sequence so that for all $n \in \omega $ , $A_n \subseteq X$ . Assume one of the following three settings:

  1. (1) $|A_n | \leq |{}^{<\delta }\lambda |$ for all $n \in \omega $ .

  2. (2) $|A_n | \leq |{}^\delta \lambda |$ for all $n \in \omega $ .

  3. (3) $|A_n | \leq |[\lambda ]^\delta |$ for all $n \in \omega .$

Then, respectively, the following hold:

  1. (1) $|\bigcup _{n \in \omega } A_n| \leq |{}^{<\delta }\lambda |$ .

  2. (2) $|\bigcup _{n \in \omega } A_n| \leq |{}^\delta \lambda |$ .

  3. (3) $|\bigcup _{n \in \omega } A_n| \leq |[\lambda ]^{\delta }|$ .

Proof. The argument is similar to the proof of Theorem 3.10 using Theorem 3.8.

Woodin defined an extension of $\mathsf {AD}$ called $\mathsf {AD}^+$ which includes (1) $\mathsf {DC}_{\mathbb {R}}$ , (2) all sets of reals are $\infty $ -Borel, and (3) ordinal determinacy. (For every $\lambda < \Theta $ , continuous function $\pi : {}^\omega \lambda \rightarrow \mathbb {R}$ , and $A \subseteq \mathbb {R}$ , the game on $\lambda $ with payoff $\pi ^{-1}[A]$ is determined.) It is open whether $\mathsf {AD}$ and $\mathsf {AD}^+$ are equivalent. Basic information about aspects of $\mathsf {AD}^+$ can be found in [Reference Caicedo and Ketchersid3, Reference Chan6, Reference Ketchersid18, Reference Larson20].

Fact 3.12. (Woodin) Suppose $\mathsf {AD}^+$ and $V = L({\mathscr {P}(\mathbb {R})})$ . Then either $\mathsf {AD}_{\mathbb {R}}$ holds or there is a set of ordinals J so that $V = L(J,\mathbb {R})$ .

Fact 3.13. If $\mathsf {AD}^+$ , $\neg \mathsf {AD}_{\mathbb {R}}$ , and $V = L({\mathscr {P}(\mathbb {R})})$ , then $\Theta $ is regular.

Proof. By Fact 3.12, there is a set of ordinals J so that $V = L(J,\mathbb {R})$ . All sets in $L(J,\mathbb {R})$ are ordinal definable from J and an $r \in \mathbb {R}$ . For each $r \in \mathbb {R}$ and $\alpha < \Theta $ , if there is an $\mathrm {OD}_{\{J,r\}}$ surjection $\varpi : \mathbb {R} \rightarrow \alpha $ , then let $\varpi _{\alpha ,r} : \mathbb {R} \rightarrow \alpha $ be the least such surjection according to the canonical wellordering of $\mathrm {OD}_{\{J,r\}}$ . For each $\alpha < \Theta $ , let $\pi _\alpha : \mathbb {R} \rightarrow \alpha $ be defined by

$$ \begin{align*}\pi_\alpha(x) = \begin{cases} \varpi_{x^{[0]}}(x^{[1]}) & \quad \text{if there is an } \mathrm{OD}_{\{J,x^{[0]}\}} \text{ surjection of } \mathbb{R} \text{ onto } \alpha \\ 0 & \quad \text{otherwise}. \end{cases}\end{align*} $$

$\pi _\alpha $ is a surjection. Thus a sequence $\langle \pi _\alpha : \alpha < \Theta \rangle $ has been defined so that $\pi _\alpha : \mathbb {R} \rightarrow \alpha $ is a surjection for each $\alpha < \Theta $ . Now suppose ${\mathrm {cof}}(\Theta ) < \Theta $ . Let $\tau : \mathbb {R} \rightarrow {\mathrm {cof}}(\Theta )$ be a surjection. Define $\sigma : \mathbb {R} \rightarrow \Theta $ by $\sigma (x) = \pi _{\tau (x^{[0]})}(x^{[1]})$ . $\sigma $ is a surjection onto $\Theta ,$ which is impossible.

Let $1 \leq n < \omega $ and $A \subseteq \mathbb {R}^n$ (again $\mathbb {R}$ refers to ${{}^\omega \omega }$ ). A is Suslin if and only if there is an ordinal $\lambda $ and a tree $T \subseteq \omega ^n \times \lambda $ so that $A = \{(x_1, \ldots , x_n) \in \mathbb {R}^n : (\exists f \in {}^\omega \lambda )((x_1, \ldots ,x_n,f) \in [T]\}$ . $A \subseteq \mathbb {R}^n$ is coSuslin if and only if $\mathbb {R}^n \setminus A$ is Suslin.

Fact 3.14. (Woodin) Assume $\mathsf {AD}^+$ and $\mathsf {AD}_{\mathbb {R}}$ . All sets of reals are Suslin.

A transitive set M is said to be Suslin and coSuslin if and only if there is a surjection $\pi : \mathbb {R} \rightarrow M$ so that the equivalence relation $E_\pi \subseteq \mathbb {R} \times \mathbb {R}$ on $\mathbb {R}$ and the relation $F_\pi \subseteq \mathbb {R} \times \mathbb {R}$ defined below are Suslin and coSuslin:

$$ \begin{align*}x \ E_\pi \ y \Leftrightarrow \pi(x) = \pi(y) \ \ \ \text{ and } \ \ \ (x,y) \in F_\pi \Leftrightarrow \pi(x) \in \pi(y).\end{align*} $$

Note that M is in bijection with . Let be defined by $([x]_{E_\pi }, [y]_{E_\pi }) \in \tilde F_\pi $ if and only if $(x,y) \in F_\pi $ . Then $(M,\in )$ is $\in $ -isomorphic to . In other words, M is Suslin and CoSuslin if it has a natural coding on $\mathbb {R}$ which is Suslin and coSuslin.

Let $\mathcal {S}$ be the union of the collection of all transitive sets which are Suslin and coSuslin. $(\mathcal {S},\in )$ is a $\in $ -structure. In general, one says a set X is Suslin and coSuslin if and only if $X \in \mathcal {S}$ .

Woodin showed that $\mathsf {AD}^+$ implies the following reflection property.

Fact 3.15. (Woodin [Reference Steel and Trang27]) ( $\Sigma _1$ -reflection into Suslin and coSuslin) Assume $\mathsf {AD}^+$ and $V = L({\mathscr {P}(\mathbb {R})})$ . $\mathcal {S} \prec _{\Sigma _1} (V,\in )$ . (That is, $\mathcal {S}$ is a $\Sigma _1$ -elementary substructure of the universe V.)

Theorem 3.16. Assume $\mathsf {AD}^+$ . Let X be a surjective image of $\mathbb {R}$ . Let $\langle A_\alpha : \alpha < {\omega _1}\rangle $ be a sequence so that for all $\alpha < {\omega _1}$ , $A_\alpha \subseteq X$ . Let $\delta $ and $\lambda $ be cardinals so that ${\omega _1} \leq \delta \leq \lambda < \Theta $ . Assume one of the following three settings:

  1. (1) ${\mathrm {cof}}(\delta ) \geq {\omega _1}$ and for all $\alpha < {\omega _1}$ , $|A_\alpha | \leq |{}^{<\delta }\lambda |$ .

  2. (2) For all $\alpha < {\omega _1}$ , $|A_\alpha | \leq |{}^\delta \lambda |$ .

  3. (3) ${\mathrm {cof}}(\lambda ) \geq {\omega _1}$ and for all $\alpha < {\omega _1}$ , $|A_\alpha | \leq |[\lambda ]^\delta |$ .

Then, respectively, the following hold:

  1. (1) $|\bigcup _{\alpha < {\omega _1}} A_\alpha | \leq |{}^{<\delta }\lambda |$ .

  2. (2) $|\bigcup _{\alpha < {\omega _1}} A_\alpha | \leq |{}^\delta \lambda |$ .

  3. (3) $|\bigcup _{\alpha < {\omega _1}} A_\alpha | \leq |[\lambda ]^\delta |$ .

Proof. Consider the setting of $(1)$ . Let $\varsigma : \mathbb {R} \rightarrow X$ be a surjection. Define an equivalence relation E on $\mathbb {R}$ by $x \ E \ y$ if and only if $\varsigma (x) = \varsigma (y)$ . Note that X is in bijection with . For each $\alpha < {\omega _1}$ , let $K_\alpha = \varsigma ^{-1}[A_\alpha ]$ and $E_\alpha = E \upharpoonright K_\alpha $ . Then and $A_\alpha $ is in bijection with . Injections of $A_\alpha $ into ${}^{<\delta }\lambda $ induce injections of into ${}^{<\delta }\lambda $ . Let be defined by $\pi (x) = [x]_E$ . Let $\varpi : \mathbb {R} \rightarrow {\mathscr {P}(\lambda )}$ be a surjection given by Fact 3.7. Then injections between and $[\lambda ]^{<\delta }$ can be coded by sets of reals through the coding $B \mapsto T_B$ described above. This shows that X and $\langle A_\alpha : \alpha < {\omega _1}\rangle $ with the property stated in setting (1) are in bijection with objects and with the properties in setting (1), which belong to $L({\mathscr {P}(\mathbb {R})})$ . It suffices to prove the theorem in $L({\mathscr {P}(\mathbb {R})})$ .

With this discussion in mind, one will now assume $\mathsf {AD}^+$ , $V = L({\mathscr {P}(\mathbb {R})})$ , and that X and $\langle A_\alpha : \alpha < {\omega _1}\rangle $ belong to $L({\mathscr {P}(\mathbb {R})})$ with the properties stated in (1). If ${\mathrm {cof}}(\Theta )> {\omega _1}$ , then the result follows from Theorem 3.10. Suppose ${\mathrm {cof}}(\Theta ) \leq {\omega _1}$ . Thus $\Theta $ is singular and hence $\mathsf {AD}_{\mathbb {R}}$ holds by Fact 3.13. Assume for the sake of contradiction that there is a set X and a sequence $\langle A_\alpha : \alpha < {\omega _1}\rangle $ satisfying (1) and $\neg (|\bigcup _{\alpha < {\omega _1}} A_\alpha | \leq |{}^{<\delta }\lambda |)$ . Let $Y = \bigcup _{\alpha < {\omega _1}} A_\alpha $ and thus $\neg (|Y| \leq |{}^{<\delta }\lambda |)$ . Since all sets of reals are Suslin and coSuslin by Fact 3.14 since $\mathsf {AD}^+$ and $\mathsf {AD}_{\mathbb {R}}$ hold, the sets Y, $\delta $ , and $\lambda $ are Suslin and coSuslin and hence belong to $\mathcal {S}$ .

Let $\psi $ be the following sentence with $\delta $ , $\lambda $ , and Y as a parameter: $\delta \leq \lambda < \dot \Theta $ and there exists a sequence $\langle \tilde A_\alpha : \alpha < {\omega _1}\rangle $ so that $Y = \bigcup _{\alpha < {\omega _1}} \tilde A_\alpha $ and for all $\alpha < {\omega _1}$ , $|A_\alpha | \leq |{}^{<\delta }\lambda |$ . ( $\dot \Theta $ is an abbreviation for the ordinal defined as the supremum of the ordinals which are surjective images of $\mathbb {R}$ .) Let $\mathfrak {T}$ be some sufficiently strong finite fragment of $\mathsf {ZF}$ . Let $\varphi $ be the following $\Sigma _1$ -sentence with Y, $\delta $ , $\lambda $ , and $\mathbb {R}$ as parameters: There exists a transitive set $M \models \mathfrak {T} + \mathsf {AD}$ so that $\mathbb {R} \subseteq M$ and $M \models \psi $ . Let $\preceq $ be a prewellordering of length $\lambda $ whose associated norm was used to define the surjection $\varpi : \mathbb {R} \rightarrow {\mathscr {P}(\lambda ),}$ which appears in the coding described before Theorem 3.8. Since $L({\mathscr {P}(\mathbb {R})}) \models $ $\mathfrak {T}$ , $\mathsf {AD}$ , and $\psi $ ” and using reflection on the hierarchy $\langle L_\alpha ({\mathscr {P}(\mathbb {R})}) : \alpha < \mathrm {ON}\rangle $ , there is an ordinal $\alpha \geq \Theta $ such that $L_\alpha ({\mathscr {P}(\mathbb {R})}) \models $ $\mathfrak {T}$ , $\mathsf {AD}$ , and $\psi $ ”. Thus $L({\mathscr {P}(\mathbb {R})}) \models \varphi $ as witnessed by $L_\alpha ({\mathscr {P}(\mathbb {R})})$ . By $\Sigma _1$ -reflection into Suslin and coSuslin (Fact 3.15), $\mathcal {S} \models \varphi $ . Let $M \in \mathcal {S}$ be a transitive set containing $\mathbb {R}$ so that $M \models \psi $ . Let $\langle \tilde A_\alpha : \alpha < {\omega _1}\rangle $ with $Y = \bigcup _{\alpha < {\omega _1}} \tilde A_\alpha $ witness the existential quantifier in $\psi $ . Since for each $\alpha < {\omega _1}$ , $M \models |\tilde A_\alpha | \leq |{}^{<\delta }\lambda |$ , $\mathbb {R} \subseteq M$ , satisfies $\mathsf {AD}$ , and has the prewellordering $\preceq $ used to code injections of subsets of Y into ${}^{<\delta }\lambda $ , there is some $B \in {\mathscr {P}(\mathbb {R})} \cap M$ so that $T_B$ codes the graph of an injection of $\tilde A_\alpha $ into ${}^{<\delta }\lambda $ . Since $M \in \mathcal {S}$ implies M is a surjective image of $\mathbb {R}$ , $\sup \{\mathrm {rk}_L(B) : B \in {\mathscr {P}(\mathbb {R})} \cap M\} < \Theta ^V$ . In the real world, let $Z \in {\mathscr {P}(\mathbb {R})}$ be such that $\mathrm {rk}_L(Z) \geq \sup \{\mathrm {rk}_L(B) : B \in {\mathscr {P}(\mathbb {R})} \cap M\}$ . Note that for all $\alpha < {\omega _1}$ , there is an $r \in \mathbb {R}$ so that $T_{\Xi _r^{-1}[Z]}$ codes the graph of an injection of $\tilde A_\alpha $ into $[\lambda ]^{<\delta }$ . Applying Theorem 3.9 in the real world to $\langle \tilde A_\alpha : \alpha < {\omega _1}\rangle $ , one has that $|Y| = |\bigcup _{\alpha < {\omega _1}} \tilde A_\alpha | \leq |{}^{<\delta }\lambda |$ . This contradicts the assumption that $\neg (|Y| \leq |{}^{<\delta }\lambda |)$ .

Theorem 3.17. Assume $\mathsf {AD}^+$ . Suppose X is a surjective image of $\mathbb {R}$ . Let $1 \leq \delta < \Theta $ and $\omega \leq \lambda < \Theta $ . Let $\langle A_n : n \in \omega \rangle $ be a sequence so that for all $n \in \omega $ , $A_n \subseteq X$ . Assume one of the following three settings:

  1. (1) $|A_n | \leq |{}^{<\delta }\lambda |$ for all $n \in \omega $ .

  2. (2) $|A_n | \leq |{}^\delta \lambda |$ for all $n \in \omega $ .

  3. (3) $|A_n | \leq |[\lambda ]^\delta |$ for all $n \in \omega .$

Then, respectively, the following hold:

  1. (1) $|\bigcup _{n \in \omega } A_n| \leq |{}^{<\delta }\lambda |$ .

  2. (2) $|\bigcup _{n \in \omega } A_n| \leq |{}^\delta \lambda |$ .

  3. (3) $|\bigcup _{n \in \omega } A_n| \leq |[\lambda ]^{\delta }|$ .

Proof. The proof follows the template of the proof of Theorem 3.16 using Theorem 3.8.

Theorem 3.18. Assume $\mathsf {AD}^+$ (or $\mathsf {AD}$ , $\mathsf {DC}_{\mathbb {R}}$ , and ${\mathrm {cof}}(\Theta )> {\omega _1}$ ). If $\langle A_\alpha : \alpha < {\omega _1}\rangle $ is a sequence such that $\bigcup _{\alpha < {\omega _1}} A_\alpha = [\omega _2]^{<\omega _2}$ , then there is an $\alpha < {\omega _1}$ so that $\neg (|A_\alpha | \leq |[\omega _2]^{\omega _1}|)$ .

Proof. Suppose $\langle A_\alpha : \alpha < {\omega _1}\rangle $ is a sequence such that $[\omega _2]^{<\omega _2} = \bigcup _{\alpha < {\omega _1}} A_\alpha $ . Suppose for the sake of contradiction that for all $\alpha < {\omega _1}$ , $|A_\alpha | \leq |[\omega _2]^{{\omega _1}}|$ . By Theorem 3.16, $|[\omega _2]^{<\omega _2}| \leq |[\omega _2]^{{\omega _1}}|$ which violates Fact 2.14.

Theorem 3.18 is regarded as partial evidence that $[\omega _2]^{<\omega _2}$ is ${\omega _1}$ -regular which means for any $\langle A_\alpha : \alpha < {\omega _1}\rangle $ such that $\bigcup _{\alpha < {\omega _1}} A_\alpha = [\omega _2]^{<\omega _2}$ , there is an $\alpha < {\omega _1}$ so that $|A_\alpha | = |[\omega _2]^{<\omega _2}|$ . This conjecture has recently been solved by the authors. The authors in [Reference Chan, Jackson and Trang8] showed that under $\mathsf {AD}$ , $[\omega _2]^{<\omega _2}$ has ${\omega _1}$ -regular cardinality. However, it is still not known if ${\mathscr {P}(\omega _2)}$ is ${\omega _1}$ -regular or even $2$ -regular. The following is some evidence.

Theorem 3.19. Assume $\mathsf {AD}^+$ (or $\mathsf {AD}$ , $\mathsf {DC}_{\mathbb {R}}$ , and ${\mathrm {cof}}(\Theta )> {\omega _1}$ ). If $\langle A_\alpha : \alpha < {\omega _1}\rangle $ is a sequence such that $\bigcup _{\alpha < {\omega _1}} A_\alpha = {\mathscr {P}(\omega _2)}$ , then there is an $\alpha < {\omega _1}$ so that $\neg (|A_\alpha | \leq |[\omega _2]^{<\omega _2}|)$ .

Proof. Suppose $\langle A_\alpha : \alpha < {\omega _1}\rangle $ is a sequence such that ${\mathscr {P}(\omega _2)} = \bigcup _{\alpha < {\omega _1}} A_\alpha $ . Suppose for the sake of contradiction that for all $\alpha < {\omega _1}$ , $|A_\alpha | \leq |[\omega _2]^{<\omega _2}|$ . By Theorem 3.16, $|{\mathscr {P}(\omega _2)}| \leq |[\omega _2]^{<\omega _2}|,$ which violates Fact 2.10.

Since under $\mathsf {AD}$ , $\omega _3$ is singular with ${\mathrm {cof}}(\omega _3) = \omega _2$ , Fact 2.9 cannot be used to show $[\omega _3]^{<\omega _3}$ or even $[\omega _3]^{\omega _2}$ have smaller cardinality than ${\mathscr {P}(\omega _3)}$ . However [Reference Chan4] shows that $|[\omega _3]^{\omega _2}| < |[\omega _3]^{<\omega _3}| \leq |{\mathscr {P}(\omega _3)}|$ under $\mathsf {AD}^+$ by the following result.

Fact 3.20. [Reference Chan4] Assume $\mathsf {AD}^+$ .

  1. (1) (ABCD Conjecture) Let $\alpha $ , $\beta $ , $\gamma $ , and $\delta $ be cardinals such that $\omega \leq \alpha \leq \beta < \Theta $ and $\omega \leq \gamma \leq \delta < \Theta $ . $|{}^\alpha \beta | \leq |{}^\gamma \delta |$ if and only if $\alpha \leq \gamma $ and $\beta \leq \delta $ .

  2. (2) If $\kappa < \Theta $ is a cardinal and $\epsilon < \kappa $ , then $|{}^\epsilon \kappa | < |{}^{<\kappa }\kappa |$ .

It is still open if $|[\omega _3]^{<\omega _3}| < |{\mathscr {P}(\omega _3)}|$ . The following result implies that if one decomposes $[\omega _3]^{<\omega _3}$ or ${\mathscr {P}(\omega _3)}$ into ${\omega _1}$ -many pieces $\langle A_\alpha : \alpha < {\omega _1}\rangle $ . Then at least one piece $A_\alpha $ does not inject into $[\omega _3]^{\omega _2}$ .

Theorem 3.21. Assume $\mathsf {AD}^+$ (or $\mathsf {AD}$ , $\mathsf {DC}_{\mathbb {R}}$ , and ${\mathrm {cof}}(\Theta )> {\omega _1}$ ).

  1. (1) If ${\omega _1} \leq \kappa < \Theta $ is a regular cardinal and $\langle A_\alpha : \alpha < {\omega _1}\rangle $ is a sequence such that $\bigcup _{\alpha < {\omega _1}} A_\alpha = {\mathscr {P}(\kappa )}$ , then there is an $\alpha < {\omega _1}$ so that $\neg (|A_\alpha | \leq |[\kappa ]^{<\kappa }|)$ .

  2. (2) If ${\omega _1} \leq \epsilon < \kappa < \Theta $ and $\langle A_\alpha : \alpha < {\omega _1}\rangle $ is a sequence such that $\bigcup _{\alpha < {\omega _1}} A_\alpha = {}^{<\kappa }\kappa $ , then there is an $\alpha < {\omega _1}$ so that $\neg (|A_\alpha | \leq |{}^\epsilon \kappa |)$ .

  3. (3) If ${\omega _1} \leq \epsilon < \kappa < \Theta $ and $\langle A_\alpha : \alpha < {\omega _1}\rangle $ is a sequence such that $\bigcup _{\alpha < {\omega _1}} A_\alpha = {\mathscr {P}(\kappa )}$ , then there is an $\alpha < {\omega _1}$ so that $\neg (|A_\alpha | \leq |{}^\epsilon \kappa |)$ .

Proof. (1) If $|A_\alpha | \leq |[\kappa ]^{<\kappa }| = |{}^{<\kappa }\kappa |$ , then $|{\mathscr {P}(\kappa )}| = |{}^{<\kappa }\kappa |$ by Theorem 3.16. Since $\mathsf {AD}^+$ implies boldface $\mathsf {GCH}$ below $\Theta $ , this would contradict Fact 2.9.

(2) If $|A_\alpha | \leq |{}^\epsilon \kappa |$ , then $|{}^{<\kappa }\kappa | = |{}^\epsilon \kappa |$ by Theorem 3.16. This would contradict Fact 3.20.

The proof of (3) is similar.

4 Decomposition into a Suslin cardinal many pieces

This section will consider a decomposition of sets into $\kappa $ many pieces, where $\kappa $ is a Suslin cardinal. Kechris and Woodin [Reference Kechris and Woodin17] developed a more general generic coding function on Suslin cardinals (or more generally reliable ordinals). In the previous section, the well-ordered additivity of the meager ideal had a prominent role in many arguments. For $\kappa> \omega $ , there is no clear analog of this for ${}^\omega \kappa $ and its generic coding function. However, if $S \subseteq \kappa $ is a countable set, then ${}^\omega S$ is homeomorphic to $\mathbb {R}$ and thus under $\mathsf {AD}$ , the meager ideal on ${}^\omega S$ (with its usual topology) will satisfy the full well-ordered additivity. The idea will be to do an argument similar to the previous section for each countable $S \subseteq \kappa $ and then take an ultrapower by a supercompact measure on $\mathscr {P}_{\omega _1}(\kappa )$ , the set of all countable subsets of $\kappa $ . One will need to impose conditions regarding the ultrapower maps of the supercompact measure to successfully generalize these arguments. However, one will still be able establish the analog of the main result of the previous section (concerning decomposition of ${\mathscr {P}(\omega _2)} = {\mathscr {P}(\boldsymbol {\delta }^1_2)}$ into ${\omega _1} = \boldsymbol {\delta }^1_1$ many pieces) for decomposition of ${\mathscr {P}(\omega _{\omega + 2})} = {\mathscr {P}(\boldsymbol {\delta }^1_4)}$ into $\omega _{\omega + 1} = \boldsymbol {\delta }^1_3$ many pieces.

Definition 4.1. [Reference Kechris and Woodin17] An ordinal $\lambda $ is reliable if and only if there is a scale $\vec \varphi = \langle \varphi _n : n \in \omega \rangle $ on a set $W \subseteq \mathbb {R}$ such that the following holds:

  1. (1) For all $n \in \omega $ , $\varphi _n : W \rightarrow \lambda $ and $\varphi _0 : W \rightarrow \lambda $ is a surjection.

  2. (2) The relation $S_0(x,y)$ defined by $x,y \in W \wedge \varphi _0(x) \leq \varphi _0(y)$ and $S_1(x,y)$ defined by $x,y \in W \wedge \varphi _0(x) < \varphi _0(y)$ are Suslin subsets of $\mathbb {R}^2$ .

$\vec \varphi $ with the above property will be called the reliability witness for $\lambda $ .

If $\sigma \subseteq \lambda $ (which is usually countable) and $\xi \in \sigma $ , then $\sigma $ is said to be $\xi $ -honest (relative to $\vec \varphi $ ) if and only if there is a $w \in W$ so that $\varphi _0(w) = \xi $ and for all $n \in \omega $ , $\varphi _n(\xi ) \in \sigma $ . Such a $w \in W$ will be called a $\xi $ -honest witness for $\sigma $ (relative to $\vec \varphi $ ). A countable $\sigma \subseteq \lambda $ is honest (relative to $\vec \varphi $ ) if and only if for all $\xi \in \sigma $ , $\sigma $ is $\xi $ -honest.

Fact 4.2. Suppose $\lambda $ is a reliable ordinal with reliability witness $\vec \varphi ,$ which is a scale on a set $W \subseteq \mathbb {R}$ . For each $\xi < \lambda $ , there is a countable set $\sigma $ so that $\sigma $ is $\xi $ -honest relative to $\vec \varphi $ .

Proof. Let $w \in W$ so that $\varphi _0(w) = \xi ,$ which is possible since $\varphi _0 : W \rightarrow \lambda $ is surjective. Let $\sigma = \{\varphi _n(w) : n \in \omega \}$ . $\sigma $ is $\xi $ -honest with w as its $\xi $ -honest witness.

It is generally not possible to uniformly associate $\xi $ to a countable $\xi $ -honest set (relative to a reliability witness). However if $\lambda $ is a reliable ordinal of uncountable cofinality, then one can at least uniformly associate $\xi $ to an ordinal $\xi ' < \lambda $ which is $\xi $ -honest which will be sufficient for applications here.

Fact 4.3. Suppose $\lambda $ is a reliable ordinal with reliability witness $\vec \varphi $ and ${\mathrm {cof}}(\lambda )> \omega $ . For each $\xi < \lambda $ , there is a $\xi ' < \lambda $ so that for all $\gamma $ with $\xi ' \leq \gamma < \lambda $ , $\gamma $ is $\xi $ -honest relative to $\vec \varphi $ .

Proof. By Fact 4.2, there is a countable $\bar \sigma \subseteq \lambda ,$ which is $\xi $ -honest. ${\xi ' = \sup (\bar \sigma ) < \lambda }$ since ${\mathrm {cof}}(\lambda )> \omega $ . Suppose $\gamma $ is such that $\xi ' \leq \gamma < \kappa $ . Since $\bar \sigma \subseteq \gamma $ , $\gamma $ is $\xi $ -honest.

Definition 4.4. Let X be a set. Let $\mathscr {P}_{\omega _1}(X) = \{\sigma \in \mathscr {P}(X) : |\sigma | \leq \omega \}$ (which is the set of countable subsets of X). Let $\nu $ be an ultrafilter on $\mathscr {P}_{\omega _1}(X)$ . $\nu $ is a fine ultrafilter on $\mathscr {P}_{\omega _1}(X)$ if and only if for each $x \in X$ , $A_x = \{\sigma \in \mathscr {P}_{\omega _1}(X) : x \in \sigma \} \in \nu $ . $\nu $ is a normal ultrafilter on $\mathscr {P}_{\omega _1}(X)$ if and only if for every $\Phi : \mathscr {P}_{\omega _1}(X) \rightarrow \mathscr {P}_{\omega _1}(X)$ such that $\{\sigma \in \mathscr {P}_{\omega _1}(X) : \emptyset \neq \Phi (\sigma ) \subseteq \sigma \} \in \nu $ , there is an $x \in X$ so that $\{\sigma \in \mathscr {P}_{\omega _1}(X) : x \in \Phi (\sigma )\} \in \nu $ . $\nu $ is a supercompact measure on X if and only if $\nu $ is a countably complete, fine, and normal measure on $\mathscr {P}_{\omega _1}(X)$ .

Fact 4.5. (Harrington–Kechris [Reference Harrington and Kechris11]) Assume $\mathsf {AD}$ . If $\kappa $ less than or equal to a Suslin cardinal, then there is a supercompact measure on $\mathscr {P}_{\omega _1}(\kappa )$ .

(Woodin [Reference Woodin28]) Assume $\mathsf {AD}$ . If $\kappa $ is less than or equal to a Suslin cardinal, then the supercompact measure on $\mathscr {P}_{\omega _1}(\kappa )$ is unique.

Fact 4.6. Assume $\mathsf {AD}$ . Suppose $\vec \varphi $ is a sequence of norms on $W \subseteq \mathbb {R}$ which is a reliability witness for $\lambda $ . Let $\nu $ be a countably complete and fine measure on $\mathscr {P}_{\omega _1}(\lambda )$ . Let $\xi < \lambda $ . Then $\{\sigma \in \mathscr {P}_{\omega _1}(\lambda ) : \sigma \text { is } \xi \text {-honest}\} \in \nu $ .

Proof. Pick any $w \in W$ so that $\varphi _0(w) = \xi $ (which is possible since $\varphi _0$ surjects onto $\lambda $ ). By fineness of $\nu $ , $A_n = \{\sigma \in \mathscr {P}_{\omega _1}(\lambda ) : \varphi _n(w) \in \sigma \} \in \nu $ . By countably compleness of $\nu $ , $\bigcap _{n \in \omega } A_n \in \nu $ . Since $\nu $ is a filter, $\bigcap _{n \in \omega } A_n \subseteq \{\sigma \in \mathscr {P}_{\omega _1}(\lambda ) : \sigma \text { is } \xi \text {-honest}\} \in \nu $ .

Fact 4.7. Assume $\mathsf {AD}$ . Suppose $\vec \varphi $ is a sequence of norms on $W \subseteq \mathbb {R}$ is a reliability witness for $\lambda $ . Let $\nu $ be a supercompact measure on $\mathscr {P}_{\omega _1}(\lambda )$ . Then $A = \{\sigma \in \mathscr {P}_{\omega _1}(\lambda ) : \sigma \text { is honest}\} \in \nu $ .

Proof. Suppose $A \notin \nu $ . Let $\tilde A = \mathscr {P}_{\omega _1}(\lambda ) \setminus A$ . Since $\nu $ is an ultrafilter, ${\tilde A \in \nu }$ . Let $\Phi : \mathscr {P}_{\omega _1}(\lambda ) \rightarrow \mathscr {P}_{\omega _1}(\lambda )$ be defined by $\Phi (\sigma ) = \{\xi \in \sigma : \sigma \text { is not} \xi -honest\}$ . Note that for all $\sigma \in \tilde A$ , $\emptyset \neq \Phi (\sigma ) \subseteq \sigma $ . So $\tilde A \subseteq \{\sigma \in \mathscr {P}_{\omega _1}(\lambda ) : \emptyset \neq \Phi (\sigma ) \subseteq \sigma \}$ and therefore $\{\sigma \in \mathscr {P}_{\omega _1}(\lambda ) : \emptyset \neq \Phi (\sigma ) \subseteq \sigma \} \in \nu $ . By normality, there is a $\eta \in \lambda $ so that ${B = \{\sigma \in \mathscr {P}_{\omega _1}(\lambda ) : \eta \in \Phi (\sigma )\} \in \nu }$ . Pick a $w \in W$ so that $\varphi _0(w) = \eta $ . For each $n \in \omega $ , $C_n = \{\sigma \in \mathscr {P}_{\omega _1}(\lambda ) : \varphi _n(w) \in \sigma \} \in \nu $ by fineness. Then $C = \bigcap _{n \in \omega } C_n \in \nu $ by countably completeness. Then $D = B \cap C \in \nu $ . Pick any $\sigma \in D$ . w is an $\eta $ -honest witness for $\sigma $ since for all $n \in \omega $ , $\varphi _n(w) \in \sigma $ . Thus $\sigma $ is $\eta $ -honest. However, $\eta \in \Phi (\sigma )$ means that $\sigma $ is not $\eta $ -honest. This is a contradiction.

Recall the notation $x^{[n]}$ from Definition 3.1 for $x \in \mathbb {R}$ and $n \in \omega $ .

Fact 4.8. (Kechris–Woodin [Reference Kechris and Woodin17, Lemma 1.1] [Reference Jackson14, Theorem 6.1] Assume $\mathsf {AD}$ . Let $\lambda $ be a reliable ordinal with $\vec \varphi $ be a sequence of norms on a set $W \subseteq \mathbb {R}$ being a reliability witness. Then there is a Lipschitz continuous function $\mathfrak {G} : {}^\omega \lambda \rightarrow \mathbb {R}$ so that the following holds:

  1. (1) For all $n \in \omega $ and $f \in {}^\omega \lambda $ , $\mathfrak {G}(f)^{[n]} \in W$ and $\varphi _0(\mathfrak {G}(f)^{[n]}) \leq f(n)$ .

  2. (2) For all $n \in \omega $ and $f \in {}^\omega \lambda $ , if $f[\omega ]$ is $f(n)$ -honest, then $\varphi _0(\mathfrak {G}(f)^{[n]}) = f(n)$ .

Thus if $f[\omega ]$ is honest, then for all $n \in \omega $ , $\varphi _0(\mathfrak {G}(f)^{[n]}) = f(n)$ . For each $n \in \omega $ , let $\mathfrak {G}_n : {}^\omega \lambda \rightarrow W$ be defined by $\mathfrak {G}_n(f) = \mathfrak {G}(f)^{[n]}$ .

A function $\mathfrak {G}$ with the above property is called a generic coding function for $\lambda $ relative to the reliability witness $\vec \varphi $ .

Theorem 4.9 will only need the concept of $\xi $ -honest for a particular ordinal $\xi < \lambda $ and will never need full honesty. Thus one will only directly use Fact 4.6 concerning fine and countably complete measures on $\mathscr {P}_{\omega _1}(\lambda )$ rather than Fact 4.7 which involves supercompact measures on $\mathscr {P}_{\omega _1}(\lambda )$ . However, it is convenient to use the uniqueness of the supercompact measure (Fact 4.5) to uniformly find long sequences of supercompact measures on various ordinals. Theorem 4.9 will just need codes for $f(0)$ rather than all of f so the function $\mathfrak {G}_0 : {}^\omega \lambda \rightarrow W$ will be used directly rather than $\mathfrak {G}$ . The full generic coding function will be used later to analyze the ultrapower of the supercompact measure.

Again, use the notation defined before Theorem 3.8: Suppose $\pi : \mathbb {R} \rightarrow X$ . Let $\delta \leq \lambda < \Theta $ and $\varpi : \mathbb {R} \rightarrow {\mathscr {P}(\lambda )}$ . If $B \subseteq \mathbb {R}$ , let $T_B = \{(x,f) : (\exists z \in B)(x = \pi (z^{[0]}) \wedge f = \varpi (z^{[1]}))\}$ . If $A \subseteq X$ and $\Phi : A \rightarrow {}^{<\delta }\lambda $ , then there is some $B \in {\mathscr {P}(\mathbb {R})}$ so that the graph of $\Phi $ is $T_B$ .

Theorem 4.9. Assume $\mathsf {AD}$ . Let X be a surjective image of $\mathbb {R}$ . Let $\kappa $ be a reliable cardinal. Let $\kappa \leq \delta \leq \lambda < \Theta $ be a cardinal with ${\mathrm {cof}}(\delta )> \omega $ . For each $\alpha \leq \kappa $ , let $\nu _\alpha $ be the unique supercompact measure on $\mathscr {P}_{\omega _1}(\alpha )$ . Suppose one of the two cases occurs:

  1. (1) $j_{\nu _\kappa }(\delta ) = \delta $ and $j_{\nu _\kappa }(\lambda ) = \lambda $ .

  2. (2) For all $\alpha < \kappa $ , $j_{\nu _\alpha }(\delta ) = \delta $ and $j_{\nu _\alpha }(\lambda ) = \lambda $ .

Let $\langle A_\alpha : \alpha < \kappa \rangle $ be a sequence so that there exists a $Z \in {\mathscr {P}(\mathbb {R})}$ with the property that for all $\alpha \in \kappa $ , $A_\alpha \subseteq X$ , $|A_\alpha | \leq |{}^{<\delta }\lambda |$ , and there is an $r \in \mathbb {R}$ so that $T_{\Xi ^{-1}_r[Z]}$ is the graph of an injection of $A_\alpha $ into ${}^{<\delta }\lambda $ . Then $|\bigcup _{\alpha < \kappa }A_\alpha | \leq |{}^{<\delta }\lambda |$ .

Proof. Let $\vec \varphi = \langle \varphi _n : n \in \omega \rangle $ be a scale on $W \subseteq \mathbb {R}$ which serves as a reliability witness for $\kappa $ . If case (1) holds, for each $\alpha < \kappa $ , let $\xi (\alpha ) = \kappa $ . If case (2) holds, let $\xi (\alpha )$ be the least $\xi ,$ which is $\alpha $ -honest relative to $\vec \varphi $ . Regardless of the case, $j_{\nu _{\xi (\alpha )}}(\delta ) = \delta $ and $j_{\xi (\alpha )}(\lambda ) = \lambda $ for all $\alpha < \kappa $ .

Define $R \subseteq W \times \mathbb {R}$ by $R(w,r)$ if and only if $T_{\Xi _r^{-1}[Z]}$ is the graph of an injection of $A_{\varphi _0(w)}$ into ${}^{<\delta }\lambda \setminus \{\emptyset \}$ . Let $\Gamma $ be a scaled pointclass containing the Suslin relations W and $S_0$ (from Definition 4.1 for $\varphi _0$ ) and closed under $\exists ^{\mathbb {R}}$ and $\wedge $ . By applying the Moschovakis coding lemma to R, $\varphi _0$ , and $\Gamma $ , there is a relation ${\bar {R} \subseteq W \times \mathbb {R}}$ so that $\bar {R} \subseteq R$ , $\bar {R} \in \Gamma $ , and for all $\alpha < \kappa $ , there is a $w \in W$ with $\varphi _0(w) = \alpha $ and $r \in \mathbb {R}$ so that $\bar {R}(w,r)$ . Let $\tilde R \subseteq W \times \mathbb {R}$ be defined by $\tilde R(w,r)$ if and only if ${w \in W \wedge (\exists v)(S_0(v,w) \wedge S_0(w,v) \wedge \bar {R}(v,r))}$ . $\tilde R \in \Gamma $ and $\mathrm {dom}(\tilde R) = W$ . Since $\Gamma $ is a scaled pointclass, let $\Lambda : W \rightarrow \mathbb {R}$ be a uniformization with the property that for all $w \in W$ , $\tilde R(w,\Lambda (w))$ . Thus for all $w \in W$ , $R(w,\Lambda (w))$ . For all $w \in W$ , $T_{\Xi ^{-1}_{\Lambda (w)}[Z]}$ is the graph of an injection of $A_{\varphi _0(w)}$ into ${}^{<\delta }\lambda \setminus \{\emptyset \}$ . For each $w \in W$ , let $\Phi _w : A_{\varphi _0(w)} \rightarrow {}^{<\delta }\lambda \setminus \{\emptyset \}$ be the injection whose graph is $T_{\Xi ^{-1}_{\Lambda (w)}}[Z]$ .

For each $x \in \bigcup _{\alpha < \kappa } A_\alpha $ , let $\iota (x)$ be the least $\alpha < \kappa $ so that $x \in A_\alpha $ . Let $\tau : {}^{<\omega } \kappa \times \delta \times \lambda \rightarrow \lambda $ be a bijection. If $\sigma $ is a countable set and $p \in {}^{<\omega }\sigma $ , then let $N^\sigma _p = \{f \in {}^\omega \sigma : p \subseteq f\}$ . ${}^\omega \sigma $ is given the product of the discrete topology on $\sigma $ which equivalently is generated by $\{N^\sigma _p : p \in {}^{<\omega }\sigma \}$ as a basis. For any countable $\sigma $ , ${}^\omega \sigma $ is homeomorphic to ${}^\omega \omega $ and has the Baire property for its topology. For $p \in {}^{<\omega }\sigma $ and $\varphi $ a formula, $(\forall ^{*,\sigma }_p f)\varphi (f)$ abbreviates $\{f \in N^\sigma _p : \varphi (f)\}$ is comeager in $N^\sigma _p$ . For all $x \in \bigcup _{\alpha < \kappa } A_\alpha $ and $\sigma \in \mathscr {P}_{\omega _1}(\xi (\iota (x)))$ with $\iota (x) \in \sigma $ , let

$$ \begin{align*}\Upsilon^x(\sigma) &= \{\tau(p,\eta,\zeta) : p \in {}^{<\omega}\sigma \wedge (\exists \epsilon < \delta)(\forall^{*,\sigma}_{\langle \iota(x)\rangle\hat{\ }p} f)(\epsilon = \mathrm{dom}(\Phi_{\mathfrak{G}_0(f)}(x))\\& \qquad \wedge \eta < \epsilon \wedge \Phi_{\mathfrak{G}_0(f)}(x)(\eta) = \zeta)\}.\end{align*} $$

Since $\tau $ maps into $\lambda $ , one has that $\Upsilon ^x(\sigma ) \in {\mathscr {P}(\lambda )}$ . Thus for each $x \in \bigcup _{\alpha \in \kappa } A_\alpha $ , $\Upsilon ^x : \mathscr {P}_{\omega _1}(\xi (\iota (x))) \rightarrow {\mathscr {P}(\lambda )}$ . Note that the hypothesis that implicitly implies that this ultrapower is wellfounded. Define $\Upsilon (x)$ to be the set of all ordinals $\gamma $ such that there exist (equivalently, for all) functions $f : \mathscr {P}_{\omega _1}(\xi (\iota (x))) \rightarrow \mathrm {ON}$ with $[f]_{\nu _{\xi (\iota (x))}} = \gamma $ , $\{\sigma \in \mathscr {P}_{\omega _1}(\xi (\iota (x))) : f(\sigma ) \in \Upsilon ^x(\sigma )\} \in \nu _{\xi (\iota (x))}$ . (Although this ultrapower does not satisfy Łoś’ Theorem, $\Upsilon $ is intuitively defined by $\Upsilon (x) = [\Upsilon ^x]_{\nu _{\xi (\iota (x))}}$ .)

Claim 1: For all $x \in \bigcup _{\alpha < \kappa } A_\alpha $ , $\Upsilon (x) \subseteq \lambda $ .

To see Claim 1: Suppose $\gamma \in \Upsilon (x)$ and $f : \mathscr {P}_{\omega _1}(\xi (\iota (x))) \rightarrow \mathrm {ON}$ with ${[f]_{\nu _{\xi (\iota (x))}} = \gamma }$ . Thus $\{\sigma \in \mathscr {P}_{\omega _1}(\xi (\iota (x))) :f(\sigma ) \in \Upsilon ^x(\sigma ) \subseteq {\mathscr {P}(\lambda )}\} \in \nu _{\xi (\iota (x))}$ . Thus $[f]_{\nu _{\xi (\iota (x))}} < j_{\nu _{\xi (\iota (x))}}(\lambda ) = \lambda $ . Thus $\gamma < \lambda $ . This shows $\gamma \in \lambda $ . Claim 1 has been established.

Claim 2: For all $x \in \bigcup _{\alpha < \kappa } A_\alpha $ , $\Upsilon (x) \neq \emptyset $ .

To see Claim 2: Since $\xi (\iota (x))$ is an $\iota (x)$ -honest ordinal, $A = \{\sigma \in \mathscr {P}_{\omega _1}(\xi (\iota (x))) : \sigma \text { is} \iota (x)\text {-honest}\} \in \nu _{\xi (\iota (x))}$ . Pick any $\sigma \in A$ . Let $\mathsf {surj}_\sigma ^{\iota (x)} = \{f \in {}^\omega \sigma : f[\omega ] = \sigma \wedge f(0) = \iota (x)\}$ which is a comeager subset of $N^\sigma _{\langle \iota (x)\rangle }$ . For all $f \in \mathrm {surj}^{\iota (x)}_\sigma $ , $f[\omega ] = \sigma $ is $\iota (x)$ -honest or equivalently $f(0)$ -honest. By Fact 4.8, $\varphi _0(\mathfrak {G}_0(f)) = \iota (x)$ and therefore, $\Phi _{\mathfrak {G}_0(f)} : A_{\iota (x)} \rightarrow {}^{<\delta }\lambda $ . For all $\epsilon < \delta $ , let ${B_\epsilon = \{f \in \mathsf {surj}^{\iota (x)}_\sigma : \mathrm {dom}(\Phi _{\mathfrak {G}_0(f)}(x)) = \epsilon \}}$ . One has that $\mathrm {surj}^{\iota (x)}_\sigma = \bigcup _{\epsilon < \delta } B_\epsilon $ . Since well-ordered union of meager sets is meager and $\mathrm {surj}^{\iota (x)}_\sigma $ is a comeager subset of $N^{\sigma }_{\langle \iota (x)\rangle }$ , there is some $\bar {\epsilon }$ so that $B_{\bar {\epsilon }}$ is nonmeager. (Note that $\bar \epsilon> 0$ since $\Phi _{\mathfrak {G}_0(f)} : A_{\iota (x)} \rightarrow {}^{<\delta }\lambda \setminus \{\emptyset \}$ .) For each $\zeta < \lambda $ , let $C_\zeta = \{f \in B_{\bar {\epsilon }} : \Phi _{\mathfrak {G}_0(f)}(x)(0) = \zeta \}$ . $B_{\bar \epsilon } = \bigcup _{\zeta < \lambda } C_{\zeta }$ . Again since well-ordered union of meager subsets of ${}^\omega \sigma $ are meager and $B_{\bar \epsilon }$ is nonmeager, there is $\bar {\zeta }$ so that $C_{\bar {\zeta }}$ is nonmeager. Since ${}^\omega \sigma $ has the Baire property, there is a $\bar {p} \in {}^{<\omega }\sigma $ so that $B_{\bar \epsilon }$ is comeager in $N^\sigma _{\langle \iota (x)\rangle \hat {\ }p}$ . Then $\tau (\bar {p},0,\bar {\zeta }) \in \Upsilon ^x(\sigma )$ . This shows that for all $\sigma \in A$ , $\Upsilon ^x(\sigma ) \neq \emptyset $ . Let $h : A \rightarrow \lambda $ be defined by $h(\sigma ) = \min (\Upsilon ^x(\sigma ))$ . Then $[h]_{\nu _{\xi (\iota (x))}} \in \Upsilon (x)$ . This establishes Claim 2.

Claim 3: For all $x \in \bigcup _{\alpha < \kappa } A_\alpha $ and $\sigma \in \mathscr {P}_{\omega _1}(\xi (\iota (x)))$ , $|\Upsilon ^x(\sigma )| < \delta $ .

To see Claim 3: Let $B = \{p \in {}^{<\omega } \sigma : (\exists \epsilon )(\forall ^{*,\sigma }_{\langle \iota (x)\rangle \hat {\ }p} f)(\epsilon = \mathrm {dom}(\Phi _{\mathcal {G}_0(f)}(x)))\}$ . For each $p \in B$ , there is a unique $\epsilon _p < \delta $ so that $(\forall ^{*,\sigma }_{\langle \iota (x)\rangle \hat {\ }p} f)(\epsilon _p = \mathrm {dom}(\Phi _{\mathfrak {G}_0(f)}))$ . Thus $\epsilon _p$ surjects onto $K^\sigma _p = \{\tau (p,\eta ,\zeta ) : \tau (p,\eta ,\zeta ) \in \Upsilon ^x(\sigma )\}$ since if $\tau (p,\eta ,\zeta ) \in K^\sigma _p$ , then ${\eta < \epsilon _p}$ and $\zeta $ is uniquely determined from p and $\eta $ . Hence $|K^\sigma _p| \leq |\epsilon _p| < \delta $ . Since ${B \subseteq {}^{<\omega }\sigma }$ is countable, $\Upsilon ^x(\sigma ) = \bigcup _{p \in B} K^\sigma _p$ , and ${\mathrm {cof}}(\delta )> \omega $ , one has that $|\Upsilon ^x(\sigma )| < \delta $ .

Claim 4: For all $x \in \bigcup _{\alpha < \kappa } A_\alpha $ , $|\Upsilon (x)| < \delta $ and thus $\Upsilon (x) \in \mathscr {P}_{\delta }(\lambda )$ .

To see Claim 4: Suppose $\gamma \in \Upsilon (x)$ and $[f]_{\nu _{\xi (\iota (x))}} = \gamma $ . For each $\sigma \in \mathscr {P}_{\omega _1}(\xi (\iota (x)))$ , let $h_f(\sigma )$ be the ordertype of $f(\sigma )$ in $\Upsilon ^x(\sigma )$ . By Claim 3, $h_f : \mathscr {P}_{\omega _1}(\xi (\iota (x))) \rightarrow \delta $ . Let $\Sigma ^x(\gamma ) = [h_f]_{\nu _{\xi (\iota (x))}}$ and note that $\Sigma ^x(\gamma )$ is independent of the choice of representative f. Let $g^x : \mathscr {P}_{\omega _1}(\xi (\iota (x))) \rightarrow \delta $ be defined by $g^x(\sigma ) = {\mathrm {ot}}(\Upsilon ^x(\sigma ))$ . Note that $g^x(\sigma ) < \delta $ by Claim 3. Thus $\Sigma ^x(\gamma ) = [h_f]_{\nu _{\xi (\iota (x))}} < [g^x]_{\nu _{\xi (\iota (x))}} < j_{\nu _{\xi (\iota (x))}}(\delta ) = \delta $ . Thus $\Sigma ^x : \Upsilon (x) \rightarrow [g^x]_{\nu _{\xi (\iota (x))}}$ where $[g^x]_{\nu _{\xi (\iota (x))}} < \delta $ . Now suppose $\gamma _0 < \gamma _1$ and $\gamma _0,\gamma _1 \in \Upsilon (x)$ . Let $f_0$ and $f_1$ be such that $[f_0]_{\nu _{\xi (\iota (x))}} = \gamma _0$ and $[f_1]_{\nu _{\xi (\iota (x))}} = \gamma _1$ . Thus $\{\sigma \in \mathscr {P}_{\omega _1}(\xi (\iota (x))) : f_0(\sigma ) < f_1(\sigma )\} \in \nu _{\xi (\iota (x))}$ . Thus $\Sigma ^x(\gamma _0) = [h_{f_0}]_{\nu _{\xi (\iota (x))}} < [h_{f_1}]_{\nu _{\xi (\iota (x))}} = \Sigma ^x(\gamma _1)$ . Thus $\Sigma ^x : \Upsilon (x) \rightarrow [g^x]_{\nu _{\xi \iota (x)}}$ is an order-preserving map. Thus $|\Upsilon (x)| < \delta $ and hence $\Upsilon (x) \in \mathscr {P}_{\delta }(\lambda )$ . This shows Claim 4.

Define $\chi : \bigcup _{\alpha < \kappa } A_\alpha \rightarrow \kappa \times \mathscr {P}_\delta (\lambda )$ by $\chi (x) = (\iota (x),\Upsilon (x))$ .

Claim 5: $\chi : \bigcup _{\alpha < \kappa } A_\alpha \rightarrow \kappa \times \mathscr {P}_\delta (\lambda )$ is an injection.

To see Claim 5: Suppose $x_0,x_1 \in \bigcup _{\alpha < \kappa } A_\alpha $ and $x_0 \neq x_1$ . First suppose ${\iota (x_0) \neq \iota (x_1)}$ . Then $\chi (x_0) = (\iota (x_0),\Upsilon (x_0)) \neq (\iota (x_1),\Upsilon (x_1)) = \chi (x_1)$ . Now assume $\iota (x_0) = \iota (x_1)$ and let $\alpha $ be this common ordinal. Let $A = \{\sigma \in \mathscr {P}_{\omega _1}(\xi (\alpha )) : \sigma \text { is } \alpha \text {-honest}\}$ and note that $A \in \nu _{\xi (\alpha )}$ . Let $A_0$ be the set of $\sigma \in A$ so that $E^\alpha _\sigma = \{f \in \mathsf {surj}^{\alpha }_\sigma : \mathrm {dom}(\Phi _{\mathfrak {G}_0(f)}(x_0)) = \mathrm {dom}(\Phi _{\mathfrak {G}_0(f)}(x_1))\}$ is nonmeager in ${}^\omega \sigma $ . Let $A_1 = \mathsf {surj}^\alpha _{\sigma } \setminus A_0$ . Since $A = A_0 \cup A_1$ and $A \in \nu _{\xi (\alpha )}$ , exactly one of $A_0 \in \nu _{\xi (\alpha )}$ or $A_1 \in \nu _{\xi (\alpha )}$ . Suppose $A_0 \in \nu _{\xi (\alpha )}$ . Fix $\sigma \in A_0$ so $E^\alpha _\sigma $ is nonmeager. Let $F^\sigma _\epsilon = \{f \in E^\alpha _\sigma : \mathrm {dom}(\Phi _{\mathfrak {G}_0(f)}(x_0)) = \epsilon = \mathrm {dom}(\Phi _{\mathfrak {G}_0(f)}(x_1))\}$ . Since $E^\alpha _\sigma = \bigcup _{\epsilon < \delta } F^\alpha _\sigma $ and $E^\alpha _\sigma $ is nonmeager in ${}^\omega \sigma $ , let $\bar \epsilon _\sigma < \delta $ be the least $\epsilon $ so that $F^\sigma _\epsilon $ is nonmeager. Since for all $f \in F^\sigma _{\bar \epsilon _\sigma }$ , $\Phi _{\mathfrak {G}_0(f)} : A_\alpha \rightarrow {}^{<\delta }\lambda $ is an injection, $\Phi _{\mathfrak {G}_0(f)}(x_0) \neq \Phi _{\mathfrak {G}_0(f)}(x_1)$ . For each $\eta < \bar \epsilon _\sigma $ , let $H^\sigma _\eta $ be the set of $f \in F^\sigma _{\bar \epsilon _\sigma }$ so that $\eta $ is least $\eta '$ so that $\Phi _{\mathfrak {G}_0(f)}(x_0)(\eta ') \neq \Phi _{\mathfrak {G}_0(f)}(x_1)(\eta ')$ . Since $F^\sigma _{\bar \epsilon _\sigma } = \bigcup _{\eta < \bar \epsilon _\sigma } H^\sigma _\eta $ , let $\bar \eta _\sigma $ be the least $\eta $ so that $H^\sigma _{\eta }$ is nonmeager. For each pair $(\zeta _0,\zeta _1)$ of distinct ordinals in $\lambda $ , let $K^\sigma _{\zeta _0,\zeta _1}$ be the set of $f \in H^\sigma _{\bar \eta _\sigma }$ so that $\Phi _{\mathfrak {G}_0(f)}(x_0)(\bar \eta _\sigma ) = \zeta _0$ and $\Phi _{\mathfrak {G}_0(f)}(x_1)(\bar \eta _\sigma ) = \zeta _1$ . Since $H^\sigma _{\bar \eta _\sigma } = \bigcup \{K^\sigma _{\zeta _0,\zeta _1} : \zeta _0,\zeta \in \lambda \wedge \zeta _0 \neq \zeta _1\}$ , let $(\bar \zeta _0^\sigma ,\bar \zeta _1^\sigma )$ be least pair $(\zeta _0,\zeta _1)$ so that $K^\sigma _{\zeta _0,\zeta _1}$ is nonmeager. Since ${}^\omega \sigma $ has the Baire property, let $\bar {p}_\sigma $ be the least $p \in {}^{<\omega }\sigma $ (under a uniformly defined wellordering of ${}^{<\omega }\sigma )$ so that $K^\sigma _{\bar \zeta ^\sigma _0,\bar \zeta ^\sigma _1}$ is comeager in $N^\sigma _{p}$ . Then $\tau (\bar {p}_\sigma ,\bar {\eta }_\sigma ,\bar {\zeta }^\sigma _0) \in \Upsilon ^{x_0}(\sigma )$ but $\tau (\bar {p}_\sigma ,\bar {\eta }_\sigma ,\bar {\zeta }^\sigma _0) \notin \Upsilon ^{x_1}(\sigma )$ . Let $h(\sigma ) = \tau (\bar {p}_\sigma ,\bar \eta _\sigma ,\bar {\zeta }^\sigma _0)$ . Then $h(\sigma ) \in \Upsilon ^{x_0}(\sigma )$ but $h(\sigma ) \notin \Upsilon ^{x_1}(\sigma )$ for all $\sigma \in A_0$ . Then $[h]_{\nu _{\xi (\alpha )}} \in \Upsilon (x_0)$ but $[h]_{\nu _{\xi (\alpha )}} \notin \Upsilon (x_1)$ . So $\Upsilon (x_0) \neq \Upsilon (x_1)$ . Hence $\chi (x_0) = (\alpha ,\Upsilon (x_0)) \neq (\alpha ,\Upsilon (x_1)) = \chi (x_1)$ . Now suppose $A_1 \in \nu _{\xi (\alpha )}$ . Let $\sigma \in A_1$ . Then $E^\alpha _\sigma $ is meager in ${}^\omega \sigma $ . Let $I^\alpha _\sigma = \mathsf {surj}^\alpha _\sigma \setminus E^\alpha _\sigma $ , which is comeager in ${}^\omega \sigma $ . For each pair of $\epsilon _0 \neq \epsilon _1$ less than $\delta $ , let $J^\sigma _{\epsilon _0,\epsilon _1}$ be the set of $f \in I^\alpha _\sigma $ so that $\mathrm {dom}(\Phi _{\mathfrak {G}_0(f)}(x_0)) = \epsilon _0$ and $\mathrm {dom}(\Phi _{\mathfrak {G}_0(f)}(x_1)) = \epsilon _1$ . Then $I^\alpha _\sigma = \bigcup \{J^\sigma _{\epsilon _0,\epsilon _1} : \epsilon _0, \epsilon _1 < \delta \wedge \epsilon _0 \neq \epsilon _1\}$ . Let $(\bar \epsilon ^\sigma _0,\bar \epsilon ^\sigma _1)$ be the least pair $(\epsilon _0,\epsilon _1)$ with $\epsilon _0 \neq \epsilon _1$ so that $J^\sigma _{\epsilon _0,\epsilon _1}$ is nonmeager. Without loss of generality, suppose $\bar \epsilon ^\sigma _0 < \bar \epsilon ^\sigma _1$ . For each $\zeta < \lambda $ , let $Q^\sigma _\zeta = \{f \in J^\sigma _{\bar \epsilon ^\sigma _0,\bar \epsilon ^\sigma _1} : \Phi _{\mathfrak {G}_0(f)}(x_1)(\bar \epsilon _0) = \zeta \}$ . $J^\sigma _{\bar \epsilon ^\sigma _0,\bar \epsilon ^\sigma _1} = \bigcup _{\zeta < \lambda } Q^\sigma _\zeta $ . Let $\bar \zeta _\sigma $ be least $\zeta $ so that $Q^\sigma _{\zeta }$ is nonmeager. Since ${}^\omega \sigma $ has the Baire property, let $\bar {p}_\sigma $ be the least $p \in {}^{<\omega }\sigma $ so that $Q^\sigma _{\bar \zeta _\sigma }$ is comeager in $N^\sigma _p$ . Let $h(\sigma ) = \tau (\bar {p}_\sigma ,\bar \epsilon ^\sigma _0,\bar \zeta _\sigma )$ . For all $\sigma \in A_1$ , $h(\sigma ) \in \Upsilon ^{x_1}(\sigma )$ however $h(\sigma ) \notin \Upsilon ^{x_0}(\sigma )$ . Thus $[h]_{\nu _{\xi (\alpha )}} \in \Upsilon (x_1)$ and $[h]_{\nu _{\xi (\alpha )}} \notin \Upsilon (x_0)$ . So $\Upsilon (x_0) \neq \Upsilon (x_1)$ . Therefore, $\chi (x_0) = (\alpha ,\Upsilon (x_0)) \neq (\alpha ,\Upsilon (x_1)) = \chi (x_1)$ . Claim 5 has been established.

Since $|\mathscr {P}_\delta (\lambda )| = |{}^{<\delta }\lambda |$ by Fact 2.2 and $|\mathscr {P}_\delta (\lambda )| = |\kappa \times \mathscr {P}_\delta (\lambda )|$ , one has that there is an injection of $\bigcup _{\alpha < \kappa }A_\alpha $ into ${}^{<\delta }\lambda $ .

Theorem 4.10. Assume $\mathsf {AD}$ and $\mathsf {DC}_{\mathbb {R}}$ . Suppose X is a surjective image of $\mathbb {R}$ . Let $\kappa $ be a reliable cardinal. Assume ${\mathrm {cof}}(\Theta )> \kappa $ . Let $\delta $ and $\lambda $ be cardinals such that $\kappa \leq \delta \leq \lambda < \Theta $ and ${\mathrm {cof}}(\delta )> \omega $ . For each $\alpha \leq \kappa $ , let $\nu _\alpha $ be the unique supercompact measure on $\mathscr {P}_{\omega _1}(\alpha )$ . Suppose one of the two cases occurs:

  1. (1) $j_{\nu _\kappa }(\delta ) = \delta $ and $j_{\nu _\kappa }(\lambda ) = \lambda $ .

  2. (2) For all $\alpha < \kappa $ , $j_{\nu _\alpha }(\delta ) = \delta $ and $j_{\nu _\alpha }(\lambda ) = \lambda $ .

Let $\langle A_\alpha : \alpha < \kappa \rangle $ be a sequence so that for all $\alpha \in \kappa $ , $A_\alpha \subseteq X$ , and $|A_\alpha | \leq |{}^{<\delta }\lambda |$ . Then $|\bigcup _{\alpha < \kappa }A_\alpha | \leq |{}^{<\delta }\lambda |$ .

Proof. The proof follows from Theorem 4.9 in a manner similar to how Theorem 3.10 follows from Theorem 3.9.

Theorem 4.11. Assume $\mathsf {AD}^+$ . Suppose X is a surjective image of $\mathbb {R}$ . Let $\kappa $ be a reliable cardinal, which is below a Suslin cardinal. Let $\kappa \leq \delta \leq \lambda < \Theta $ be cardinals with ${\mathrm {cof}}(\delta )> \omega $ . For each $\alpha \leq \kappa $ , let $\nu _\alpha $ be the unique supercompact measure on $\mathscr {P}_{\omega _1}(\alpha )$ . Suppose one of the cases occurs:

  1. (1) $j_{\nu _\kappa }(\delta ) = \delta $ and $j_{\nu _\kappa }(\lambda ) = \lambda $ .

  2. (2) For all $\alpha < \kappa $ , $j_{\nu _\alpha }(\delta ) = \delta $ and $j_{\nu _\alpha }(\lambda ) = \lambda $ .

Let $\langle A_\alpha : \alpha < \kappa \rangle $ be a sequence so that for all $\alpha \in \kappa $ , $A_\alpha \subseteq X$ , and $|A_\alpha | \leq |{}^{<\delta }\lambda |$ . Then $|\bigcup _{\alpha < \kappa }A_\alpha | \leq |{}^{<\delta }\lambda |$ .

Proof. This result follows from Theorems 4.9 and 4.10 as in the proof of Theorem 3.16.

It is implicit in the assumption that $j_{\nu _\alpha }(\lambda ) = \lambda $ that the ultrapower is wellfounded. This is addressed in Fact 4.22. Then next few results will work toward showing $j_{\nu _\alpha }(\boldsymbol {\delta }^1_4) = \boldsymbol {\delta }^1_4$ , which is due to Becker [Reference Becker1, Theorem 4.2] One will need an explicit characterization of the supercompact measure on $\mathscr {P}_{\omega _1}(\kappa )$ when $\kappa $ is a reliable ordinal. Various constructions of a supercompact measure on $\mathscr {P}_{\omega _1}(\kappa )$ can be found in Solovay [Reference Solovay23], Harrrington–Kechris [Reference Harrington and Kechris11], and Becker [Reference Becker1]. By Woodin’s result [Reference Woodin28] concerning the uniqueness of the supercompact measure on $\mathscr {P}_{\omega _1}(\kappa )$ , they all define the same measure. Here, one will use a construction of the supercompact measure from generic codings presented in [Reference Jackson14]. However, one uses the “ordinal determinacy” clause of $\mathsf {AD}^+$ to get the necessary determinacy of certain games with moves on the ordinal. Many results below have $\mathsf {AD}^+$ as a hypothesis but had previously been proved under $\mathsf {AD}$ using the determinacy of certain real games given by [Reference Harrington and Kechris11] Harrington–Kechris. The generic coding methods seem more suitable for generalization as Becker–Jackson [Reference Becker and Jackson2] and Jackson [Reference Jackson13] showed certain cardinals (for instance, the projective ordinals $\boldsymbol {\delta }^1_n$ ) have higher degree of supercompactness (i.e., are $\boldsymbol {\delta }^2_1$ -supercompact).

Fact 4.12. Let $\kappa $ be an ordinal, $\nu $ be a supercompact measure on $\mathscr {P}_{\omega _1}(\kappa )$ , and $f : {}^{<\omega }\kappa \rightarrow \kappa $ be a function. Then $\{\sigma \in \mathscr {P}_{\omega _1}(\kappa ) : f[{}^{<\omega }\sigma ] \subseteq \sigma \} \in \nu $ .

Proof. Let $A = \{\sigma \in \mathscr {P}_{\omega _1}(\kappa ) : f[{}^{<\omega }\sigma ] \subseteq \sigma \}$ . For the sake of contradiction, suppose $A \notin \nu $ . Let $\tilde A = \mathscr {P}_{\omega _1}(\kappa ) \setminus A$ and note that $\tilde A \in \nu $ since $\nu $ is an ultrafilter. Fix a wellordering $\prec $ of ${}^{<\omega }\kappa $ . If $\sigma \in \tilde A$ , then there is a $p \in {}^{<\omega }\kappa $ so that $f(p) \notin \sigma $ . Let $p_\sigma $ be the least such p according to $\prec $ . By the countably additivity of $\nu $ , there is an $\bar {n}$ so that $B = \{\sigma \in \tilde A : |p_\sigma | = n\} \in \nu $ . If $\bar {n} = 0$ , then $p_\sigma = \emptyset $ for all $\sigma \in B$ . By fineness, $C = \{\sigma \in B : f(\emptyset ) \in \sigma \} \in \nu $ . For all $\sigma \in C$ , $f(p_\sigma ) = f(\emptyset ) \in \sigma ,$ which contradicts the definition of $p_\sigma $ . Now suppose $\bar {n}> 0$ . For each $k < \bar {n}$ , let $\Phi _k : B \rightarrow \mathscr {P}_{\omega _1}(\kappa )$ be defined by $\Phi _k(\sigma ) = \{p_\sigma (k)\}$ . For all $k < \bar {n}$ , $\{\sigma \in B : \emptyset \neq \Phi _k(\sigma ) \subseteq \sigma \} \in \nu $ . By normality, there is an $\alpha _k \in \kappa $ so that $D_k = \{\sigma \in B : \alpha _k \in \Phi _k(\sigma )\} \in \nu $ . Let $\bar {p} \in {}^{\bar {n}}\kappa $ be defined by $\bar {p}(k) = \alpha _k$ . Thus $E = \{\sigma \in B : p_\sigma = \bar {p}\} = \bigcap _{k < \bar {n}} D_k \in \nu $ by the countably completeness of $\nu $ . By fineness, $F = \{\sigma \in D : f(\bar {p}) \in \sigma \} \in \nu $ . For all $\sigma \in F$ , $f(p_\sigma ) = f(\bar {p}) \in \sigma ,$ which contradicts the definition of $p_\sigma $ . This completes the proof.

Definition 4.13. Formally a strategy on $\kappa $ is a function $\rho : {}^{<\omega }\kappa \rightarrow \kappa $ . If $\rho _0$ and $\rho _1$ are two strategies, then $\rho _0 * \rho _1 \in {}^\omega \kappa $ is defined by recursion as follows: If n is even, then $(\rho _0 * \rho _1)(n) = \rho _0(\rho _0 * \rho _1 \upharpoonright n)$ . If n is odd, then $(\rho _0 * \rho _1)(n) = \rho _1(\rho _0 * \rho _1 \upharpoonright n)$ . If $f \in {}^\omega \kappa $ , then let $\rho ^1_f$ be the strategy defined by $\rho ^1_f(2n) = f(n)$ and $\rho ^1_f(2n + 1) = 0$ for all $n \in \omega $ . If $f \in {}^\omega \kappa $ , then let $\rho ^2_f$ be the strategy defined by $\rho ^2_f(2n) = 0$ and $\rho ^2_f(2n + 1) = f(n)$ . If $f \in {}^\omega \kappa $ , let $f_{\mathrm {even}} \in {}^\omega \kappa $ and $f_{\mathrm {odd}} \in {}^\omega \kappa $ be defined by $f_{\mathrm {even}}(n) = f(2n)$ and $f_{\mathrm {odd}}(n) = f(2n + 1)$ . If $\rho $ is a strategy, then let $\Xi ^1_\rho , \Xi ^2_\rho : {}^\omega \kappa \rightarrow {}^\omega \kappa $ be defined by $\Xi ^1_\rho (f) = (\rho * \rho ^2_f)_{\mathrm {even}}$ and $\Xi ^2_\rho (f) = (\rho ^1_f * \rho )_{\mathrm {odd}}$ .

Fix a bijection $\pi ^{\kappa ,2} : \kappa \rightarrow \kappa \times \kappa $ . Let $\pi ^{\kappa ,2}_0, \pi ^{\kappa ,2}_1 : \kappa \rightarrow \kappa $ be defined by $\pi _0^{\kappa ,2}(\alpha ) = \beta $ and $\pi _1^{\kappa ,2}(\alpha ) = \gamma ,$ where $\pi ^{\kappa ,2}(\alpha ) = (\beta ,\gamma )$ . If $\rho $ is a strategy on $\kappa $ , let $\chi ^\kappa _{\rho } = \pi ^{\kappa ,2}_0 \circ \rho $ and $\tau ^\kappa _\rho = \pi ^{\kappa ,2}_1 \circ \rho $ .

Definition 4.14. Let $\kappa $ be a reliable ordinal with reliability witness $\vec \varphi ,$ which is a scale on $W \subseteq \mathbb {R}$ . Let $\rho : {}^{<\omega }\kappa \rightarrow \kappa $ be a strategy on $\kappa $ . Let $K_\rho $ be the set of $\sigma \in \mathscr {P}_{\omega _1}(\kappa )$ so that $\sigma $ is honest relative to the reliability witness $\vec \varphi $ and $\rho [{}^{<\omega }\sigma ] \subseteq \sigma $ .

Fact 4.15. Let $\kappa $ be a reliable ordinal with reliability witness $\vec \varphi ,$ which is a scale on $W \subseteq \mathbb {R}$ . Let $\rho : {}^{<\omega }\kappa \rightarrow \kappa $ be a strategy on $\kappa $ . Then $K_\rho \in \nu _\kappa $ .

Proof. This follows from Facts 4.7 and 4.12.

Generic coding can be used to define the unique supercompact measure on $\mathscr {P}_{\omega _1}(\kappa )$ when $\kappa $ is a reliable ordinal. The game will be provided next and used to show that sets of the form $K_\rho $ for strategies $\rho $ on $\kappa $ form a basis for the supercompact measure on $\mathscr {P}_{\omega _1}(\kappa )$ .

Fact 4.16. Assume $\mathsf {AD}^+$ . Let $\kappa $ be a reliable ordinal with reliability witness $\vec \varphi ,$ which is a scale on $W \subseteq \mathbb {R}$ . Let $\nu _\kappa $ be the unique supercompact measure on $\mathscr {P}_{\omega _1}(\kappa )$ . Let $A \subseteq \mathscr {P}_{\omega _1}(\kappa )$ . $A \in \nu _\kappa $ if and only if there is a strategy $\rho : {}^{<\omega }\kappa \rightarrow \kappa $ so that $K_\rho \subseteq A$ .

Proof. Fix $A \subseteq \mathscr {P}_{\omega _1}(\kappa )$ . Define the game $G_A$ on $\kappa $ as follows:

Players 1 and 2 alternate playing ordinals from $\kappa $ . Player 1 plays the ordinals $\alpha _{2n}$ and Player 2 plays the ordinals $\alpha _{2n + 1}$ for all $n \in \omega $ . Player 1 wins $G_A$ if and only if $\{\varphi _0(\mathfrak {G}_n(f)) : n \in \omega \} \in A$ . Let $\nu ^*_\kappa $ be the set of all $A \subseteq \mathscr {P}_{\omega _1}(\kappa )$ so that Player 1 has a winning strategy in $G_A$ . Let $B \subseteq {{}^\omega \omega }$ be $B = \{r \in {{}^\omega \omega } : (\forall n)(r^{[n]} \in W) \wedge \{\varphi _0(r^{[n]}) : n \in \omega \} \in A\}$ . The payoff set for $G_A$ is $\mathfrak {G}^{-1}[B]$ . Since $\mathfrak {G} : {}^\omega \kappa \rightarrow {{}^\omega \omega }$ is continuous, the “ordinal determinacy” clause of $\mathsf {AD}^+$ implies that $G_A$ is determined. It can be shown that $\nu ^*_\kappa $ is a supercompact measure on $\mathscr {P}_{\omega _1}(\kappa )$ . (Thus one can define the unique supercompact measure $\nu _\kappa $ on $\mathscr {P}_{\omega _1}(\kappa )$ to be $\nu ^*_\kappa $ .)

If there is strategy $\rho $ on $\kappa $ so that $K_\rho \subseteq A$ , then $A \in \nu _\kappa $ since $K_\rho \in \nu _\kappa $ by Fact 4.15. Now suppose $A \in \nu _\kappa = \nu _\kappa ^*$ . Let $\rho $ be a Player 1 winning strategy in $G_A$ . Let $\sigma \in K_\rho $ which means that $\sigma $ is honest and $\rho [{}^{<\omega }\sigma ] \subseteq \sigma $ . Let $g : \omega \rightarrow \sigma $ be a surjection. Let $f = \rho * \rho ^2_g$ be the run of player 1 playing the terms of g against Player 1 using $\rho $ . Since $\rho [{}^{<\omega }\sigma ] \subseteq \sigma $ and $g[\omega ] = \sigma $ , one has that $f[\omega ] = \sigma $ . Since $f[\omega ] = \sigma $ is honest, by the properties of the generic coding function (Fact 4.8), $\varphi _0(\mathfrak {G}_n(f)) = f(n)$ . Thus $\{\varphi _0(\mathfrak {G}_n(f)) : n \in \omega \} = \sigma $ . Since $\rho $ is a Player 1 winning strategy, $\sigma = \{\varphi _0(\mathfrak {G}_n(f)) : n \in \omega \} \in A$ . Since $\sigma \in K_\rho $ was arbitrary, $K_\rho \subseteq A$ .

Fact 4.17. Suppose $\kappa $ be an ordinal, $\lambda < \kappa $ , and $\nu $ is a supercompact measure on $\kappa $ . Let $\Pi : \mathscr {P}_{\omega _1}(\kappa ) \rightarrow \mathscr {P}_{\omega _1}(\lambda )$ be defined by $\Pi (\sigma ) = \sigma \cap \lambda $ . Then the Rudin–Keisler pushforward $\mu = \Pi _*\nu $ defined by $A \in \mu $ if and only if $\Pi ^{-1}[A] \in \nu $ is a supercompact measure on $\mathscr {P}_{\omega _1}(\lambda )$ .

Proof. It is straightforward to see that $\mu $ is an ultrafilter and countably complete. Suppose $\alpha \in \lambda $ . Let $A = \{\tau \in \mathscr {P}_{\omega _1}(\lambda ) : \alpha \in \tau \}$ . By the fineness of $\nu $ , $B = \{\sigma \in \mathscr {P}_{\omega _1}(\kappa ) : \alpha \in \kappa \} \in \nu $ . Note that $B = \Pi ^{-1}[A]$ . By definition $A \in \mu $ . Thus $\mu $ is fine. Let $\Phi : \mathscr {P}_{\omega _1}(\lambda ) \rightarrow \mathscr {P}_{\omega _1}(\lambda )$ be such that $C = \{\tau \in \mathscr {P}_{\omega _1}(\lambda ) : \emptyset \neq \Phi (\tau ) \subseteq \tau \} \in \mu $ . Define $\Psi : \mathscr {P}_{\omega _1}(\kappa ) \rightarrow \mathscr {P}_{\omega _1}(\kappa )$ by $\Psi (\sigma ) = \Phi (\sigma \cap \lambda )$ and note that $\Psi $ actually maps into $\mathscr {P}_{\omega _1}(\lambda )$ . Let $D = \{\sigma \in \mathscr {P}_{\omega _1}(\kappa ) : \emptyset \neq \Psi (\sigma \cap \lambda ) \subseteq \sigma \}$ . Note that $D = \Pi ^{-1}[C]$ . Thus $D \in \nu $ since $C \in \mu = \Pi _* \nu $ . By the normality of $\nu $ , there is an $\alpha \in \kappa $ so that $E = \{\sigma \in \mathscr {P}_{\omega _1}(\kappa ) : \alpha \in \Psi (\sigma )\} \in \nu $ . Note that $\alpha \in \lambda $ . Let $F = \{\tau \in \mathscr {P}_{\omega _1}(\lambda ): \alpha \in \Phi (\tau )\}$ . Note that $E = \Pi ^{-1}[F]$ and hence $F \in \mu $ . This shows that $\mu $ is normal.

Using the proof of Fact 4.17, one can provide an explicit characterization of the supercompact measure on $\mathscr {P}_{\omega _1}(\lambda )$ when $\lambda $ less than or equal to a Suslin cardinal using the generic coding on a reliable ordinal greater than or equal to $\lambda $ .

Fact 4.18. Assume $\mathsf {AD}^+$ . Let $\lambda $ be less than or equal to a Suslin cardinal and let $\kappa $ be any reliable cardinal greater than or equal to $\lambda $ . Let $\vec \varphi $ be a reliability witness for $\kappa $ . For any strategy $\rho $ on $\kappa $ , let $K^\lambda _\rho = \{\sigma \cap \lambda : \sigma \in K_\rho \}$ . For any $A \subseteq \mathscr {P}_{\omega _1}(\lambda )$ , $A \in \nu _\lambda $ if and only if there is a strategy $\rho $ on $\kappa $ so that $K^\lambda _\rho \subseteq A$ .

Proof. Let $\Pi : \mathscr {P}_{\omega _1}(\kappa ) \rightarrow \mathscr {P}_{\omega _1}(\lambda )$ be defined by $\Pi (\sigma ) = \sigma \cap \lambda $ . By Fact 4.17 and the uniqueness of the supercompact measure on $\mathscr {P}_{\omega _1}(\lambda )$ , one has that $\nu _\lambda = \Pi _* \nu _\kappa $ . Suppose $A \in \nu _\lambda $ . Then $\Pi ^{-1}[A] \in \nu _\kappa $ . By Fact 4.16, there is a strategy $\rho $ on $\kappa $ so that $K_\rho \subseteq \Pi ^{-1}[A]$ . Thus $K^\lambda _\rho = \{\sigma \cap \lambda : \sigma \in K_\rho \} = \{\Pi (\sigma ) : \sigma \in K_\rho \} \subseteq A$ . Now suppose there is a strategy $\rho $ so that $K^\lambda _\rho \subseteq A$ . Since $\Pi ^{-1}[K^\lambda _\rho ] \supseteq K_\rho $ , $\Pi ^{-1}[K^\lambda _\rho ] \in \nu _\kappa $ . So $K^\lambda _\rho \in \nu _\lambda $ . Thus $A \in \nu _\lambda $ .

The following is straightforward.

Fact 4.19. Suppose $\kappa $ is an ordinal, $|\kappa | \leq \lambda < \kappa ^+$ , and $\nu $ is a supercompact measure on $\mathscr {P}_{\omega _1}(\kappa )$ . Let $\pi : \kappa \rightarrow \lambda $ be a bijection. Let $\Pi : \mathscr {P}_{\omega _1}(\kappa ) \rightarrow \mathscr {P}_{\omega _1}(\lambda )$ be defined by $\Pi (\sigma ) = \pi [\sigma ]$ . Then the Rudin–Keisler pushforward $\mu = \Pi _*\nu $ defined by $A \in \mu $ if and only if $\Pi ^{-1}[A] \in \nu $ is a supercompact measure on $\mathscr {P}_{\omega _1}(\lambda )$ .

Fact 4.20. Assume $\mathsf {AD}$ and $\mathsf {DC}_{\mathbb {R}}$ . For any $\kappa $ less than or equal to a Suslin cardinal, let $\nu _\kappa $ denote the unique supercompact measure on $\mathscr {P}_{\omega _1}(\kappa )$ . If $\lambda < \kappa ^+$ , then $\nu _\lambda $ is Rudin–Keisler reducible to $\nu _\kappa $ .

Proof. If $\lambda < \kappa $ , then Fact 4.17 defines a supercompact measure on $\mathscr {P}_{\omega _1}(\lambda ),$ which is Rudin–Keisler reducible to $\nu _\kappa $ . By Woodin uniqueness of the supercompact measure on $\mathscr {P}_{\omega _1}(\lambda )$ , this measure must be $\nu _\lambda $ . Similarly, if $\kappa \leq \lambda < \kappa ^+$ , then Fact 4.19 defines a supercompact measure on $\mathscr {P}_{\omega _1}(\lambda ),$ which is Rudin-Keiser below $\nu _\kappa $ . Again by uniqueness, this must be $\nu _\lambda $ .

Fact 4.21. Assume $\mathsf {AD}$ . The Suslin cardinals are unbounded below their supremum.

Proof. Let $\lambda $ be the supremum of the Suslin cardinals. If $\lambda = \Theta $ , then the result is clear. Suppose $\lambda < \Theta $ but $\lambda $ is not a limit of Suslin cardinals. This means $\lambda $ is the largest Suslin cardinal. Let $S_\lambda $ be the set of $\lambda $ -Suslin sets, which is equivalently the set of all Suslin sets. First, the claim is that $S_\lambda $ is closed under $\forall ^{\mathbb {R}}$ . Suppose $S_\lambda $ is not closed under $\forall ^{\mathbb {R}}$ . There is a set $A \subseteq \mathbb {R} \times \mathbb {R}$ so that $\forall ^{\mathbb {R}} A \notin S_\lambda $ . However, the proof of the second periodicity theorem of Moschovakis [Reference Moschovakis22] shows that a Suslin representation for A can be used to create a Suslin representation for $\forall ^{\mathbb {R}} A$ . Thus $\forall ^{\mathbb {R}} A$ is a Suslin set, which does not belong to $S_\lambda ,$ which is the set of all Suslin sets. This is a contradiction. [Reference Jackson14, Lemma 3.6] states that if $S_\lambda $ is closed under $\forall ^{\mathbb {R}}$ , then $\lambda $ is a limit of Suslin cardinals. This completes the proof.

Using this explicit characterization of the supercompact measure, it will be shown next that the ultrapower ordinals below $\Theta $ by the supercompact measure on $\mathscr {P}_{\omega _1}(\kappa )$ when $\kappa $ is below a Suslin cardinal is wellfounded under $\mathsf {AD}^+$ .

Fact 4.22. Assume $\mathsf {AD}^+$ . Let $\kappa $ less than or equal to a Suslin cardinal. Let $\nu _\kappa $ be the unique supercompact measure on $\mathscr {P}_{\omega _1}(\kappa )$ . Let $(\nu _\kappa )^{L({\mathscr {P}(\mathbb {R})})}$ be the unique supercompact measure on $\mathscr {P}_{\omega _1}(\kappa )$ in $L(\mathscr {P}(\mathbb {R}))$ . Let $\lambda < \Theta $ . Then $\nu _\kappa = (\nu _\kappa )^{L({\mathscr {P}(\mathbb {R})})}$ , , and is wellfounded.

Proof. Since $\kappa $ and $\lambda $ are less than $\Theta $ , there are surjections $\pi _0 : \mathbb {R} \rightarrow \kappa $ and $\pi _1 : \mathbb {R} \rightarrow \lambda $ . Thus $\pi _2 : \mathbb {R} \rightarrow \mathscr {P}_{\omega _1}(\kappa )$ defined by $\pi _2(r) = \{\pi _0(r^{[n]}) : n \in \omega \}$ is a surjection. For each $A \subseteq \mathbb {R}$ , let $C_A = \{\pi _2(r) : r \in A\}$ . For any $X \subseteq \mathscr {P}_{\omega _1}(\kappa )$ , there is an $A \in {\mathscr {P}(\mathbb {R})}$ so that $C_A = X$ . Let $\pi _3 : \mathbb {R} \rightarrow \mathscr {P}_{\omega _1}(\kappa ) \times \lambda $ be defined by $\pi _3(r) = (\pi _2(r^{[0]}),\pi _1(r^{[1]}))$ . $\pi _3$ is a surjection. For any $A \in {\mathscr {P}(\mathbb {R})}$ , let $D_A = \{\pi _3(r) : r \in A\}$ . Thus for any $f : \mathscr {P}_{\omega _1}(\kappa ) \rightarrow \lambda $ , there is an $A \in {\mathscr {P}(\mathbb {R})}$ so that $D_A$ is the graph of f. The prewellorderings corresponding to $\pi _0$ and $\pi _1$ are subsets of $\mathbb {R}$ . Thus $L(\mathscr {P}(\mathbb {R}))$ can recover $C_A$ and $D_A$ from $A \in {\mathscr {P}(\mathbb {R})}$ . This shows that $\mathscr {P}_{\omega _1}(\kappa ) = (\mathscr {P}_{\omega _1}(\kappa ))^{L({\mathscr {P}(\mathbb {R})})}$ and $\prod _{\mathscr {P}_{\omega _1}(\kappa )}\lambda = \left (\prod _{\mathscr {P}_{\omega _1}(\kappa )}\lambda \right )^{L({\mathscr {P}(\mathbb {R})})}$ .

The real world and $L({\mathscr {P}(\mathbb {R})})$ have the same Suslin cardinals since the real world and $L({\mathscr {P}(\mathbb {R})})$ have the same trees on ordinals below $\Theta = \Theta ^{L({\mathscr {P}(\mathbb {R})})}$ using the Moschovakis coding lemma. Note that since $\kappa $ is less than or equal to a Suslin cardinal in the real world, $\kappa $ is still less than or equal to a Suslin cardinal in $L({\mathscr {P}(\mathbb {R})})$ . Since the Suslin cardinals are unbounded below the supremum of the Suslin cardinals by Fact 4.21, there is a reliable ordinal (even a Suslin cardinal) $\bar \kappa \geq \kappa $ . Since $\bar \kappa $ is a reliable ordinal, fix a reliability witness $\vec \varphi $ on $W \subseteq \mathbb {R}$ . Since $\vec \varphi = \langle \varphi _n : n \in \omega \rangle $ is a scale, $\vec \varphi \in L({\mathscr {P}(\mathbb {R})})$ . For any strategy $\rho $ on $\kappa $ , let $K_\rho $ be the set of $\sigma \in \mathscr {P}_\omega (\bar \kappa )$ such that $\rho [{}^{<\omega }\sigma ] \subseteq \sigma $ and $\sigma $ is honest relative to $\vec \varphi $ . Let $K^\kappa _\rho = \{\sigma \cap \kappa : \sigma \in K_\rho \}$ . By Fact 4.18, $A \in \nu _\kappa $ if and only if there is a strategy $\tau $ on $\bar \kappa $ so that $K^\kappa _\tau \subseteq A$ . Strategies on $\bar \kappa $ are essentially subsets of $\bar \kappa $ . By using the Moschovakis coding lemma applied in $L({\mathscr {P}(\mathbb {R})})$ using a surjection of $\mathbb {R}$ onto $\bar \kappa $ in $L({\mathscr {P}(\mathbb {R})})$ (for instance, $\varphi _0$ ), one can show that the real world and $L({\mathscr {P}(\mathbb {R})})$ have the same set of strategies on $\bar \kappa $ . Note also that for any strategy $\rho $ on $\bar \kappa $ , $K^\kappa _\rho = (K^\kappa _\rho )^{L({\mathscr {P}(\mathbb {R})})}$ since the notion of honesty is absolute. Using the explicit definition of $\nu _\kappa $ (having sets of the form $K^\kappa _\rho $ as a basis) applied in the real world or $L({\mathscr {P}(\mathbb {R})})$ , one has that $\nu _\kappa = (\nu _\kappa )^{L({\mathscr {P}(\mathbb {R})})}$ . This with the previous observation that $\prod _{\mathscr {P}_{\omega _1}(\kappa )}\lambda = \left (\prod _{\mathscr {P}_{\omega _1}(\kappa )}\lambda \right )^{L({\mathscr {P}(\mathbb {R})})}$ implies that .

Since $\mathsf {AD}^+$ holds in the real world, $L({\mathscr {P}(\mathbb {R})}) \models \mathsf {AD}^+$ . By the above, is wellfounded if and only if is wellfounded. So work inside $L({\mathscr {P}(\mathbb {R})})$ and assume for the sake of contradiction that there is some $\kappa $ less than or equal to a Suslin cardinal and ordinal $\lambda < \Theta $ so that is illfounded. For each $\alpha \leq \Theta $ , let $\mathcal {W}_\alpha $ be the set of reals of Wadge rank less than $\alpha $ . Let $\varphi $ be the sentence “there exist ordinals $\alpha $ and $\beta $ so that ”. By the reflection theorem and since ${\mathscr {P}(\mathbb {R})} = \mathcal {W}_\Theta $ , there is some $\alpha $ so that . Thus $L({\mathscr {P}(\mathbb {R})}) \models \varphi $ with witnesses $\alpha $ as above and $\beta = \Theta $ . By the $\Sigma _1$ -reflection into Suslin–coSuslin (Fact 3.15), $S \prec _{\Sigma _1} L({\mathscr {P}(\mathbb {R})})$ . There exist $\alpha < \mathcal {S}$ and $\beta \in \mathcal {S}$ so that

Since $\alpha ,\beta < \Theta $ , $L_\alpha (\mathcal {W}_\beta )$ is a surjective image of $\mathbb {R}$ . Working in $L({\mathscr {P}(\mathbb {R})}) \models \mathsf {DC}_{\mathbb {R}}$ , one can find $\langle f_n : n \in \omega \rangle $ so that $f_n \in L_\alpha (\mathcal {W}_\beta )$ , $f_n : \mathscr {P}_{\omega _1}(\kappa ) \rightarrow \lambda $ , and $L_\alpha (\mathcal {W}_\beta ) \models [f_{n + 1}]_{\nu _\kappa } < [f_n]_{\nu _\kappa }$ for each $n \in \omega $ . For each $n \in \omega $ , $L_\alpha (\mathcal {W}_\beta ) \models A_n = \{\sigma \in \mathscr {P}_{\omega _1}(\kappa ) : f_{n + 1}(\sigma ) < f_n(\sigma )\} \in \nu _\kappa $ . Note $L_\alpha (\mathcal {W}_\beta ) \models \kappa $ is less than or equal to a Suslin cardinal. Thus $L_\alpha (\mathcal {W}_\beta )$ has a reliable ordinal $\bar \kappa \geq \kappa $ . Pick a reliability witness $\vec \varphi $ for $\bar \kappa $ in $L_\alpha (\mathcal {W}_\beta )$ and note that it is a reliability witness for $\bar \kappa $ in $L({\mathscr {P}(\mathbb {R})})$ . For any strategy $\rho $ on $\bar \kappa $ , define $K^\kappa _\rho $ relative to this reliability witness $\vec \varphi $ . By applying the explicit definition of the supercompact measure on $\kappa $ within $L_\alpha (\mathcal {W}_\beta )$ , for each $n \in \omega $ , there is a strategy $\rho $ on $\bar \kappa $ so that $K^\kappa _\rho \subseteq A_n$ . Again since there is surjection of $\mathbb {R}$ onto $L_\alpha (\mathcal {W}_\beta )$ in $L({\mathscr {P}(\mathbb {R})})$ , one can use $\mathsf {AC}^{\mathbb {R}}_\omega $ in $L({\mathscr {P}(\mathbb {R})})$ to find a sequence $\langle \rho _n : n \in \omega \rangle $ so that for each $n \in \omega $ , $\rho _n \in L_\alpha (\mathcal {W}_\beta )$ is a strategy on $\bar \kappa $ , and $K^\kappa _{\rho _n} \subseteq A_n$ . Note for all $n \in \omega $ , $K^\kappa _{\rho _n} \in \nu _\kappa $ . Since $L({\mathscr {P}(\mathbb {R})}) \models \nu _\kappa $ is countably compete, $\bigcap _{n \in \omega } K^\kappa _{\rho _n} \neq \emptyset $ . Let $\sigma \in \bigcap _{n \in \omega } K^\kappa _{\rho _n} \subseteq \bigcap _{n \in \omega } A_n$ . Then in $L({\mathscr {P}(\mathbb {R})})$ , $\langle f_n(\sigma ) : n \in \omega \rangle $ is an infinite descending sequence of ordinals below $\lambda $ . This is a contradiction.

Fact 4.23. (Almost everywhere honest-enumeration uniformization) Assume $\mathsf {AD}^+$ . Let $\kappa $ be a reliable ordinal with reliability witness $\vec \varphi ,$ which is a scale on a set $W \subseteq \mathbb {R}$ . Let $R \subseteq \mathscr {P}_{\omega _1}(\kappa ) \times {{}^\omega \omega }$ be such that $\mathrm {dom}(R) = \mathscr {P}_{\omega _1}(\kappa )$ . There is a strategy $\rho $ on $\kappa $ with the following properties:

  1. (1) For all $s \in {}^{<\omega }\kappa $ with $|s|$ odd, $\tau ^\kappa _\rho (s) \in \omega $ .

  2. (2) For all $f \in {}^\omega \kappa $ such that $f[\omega ] \in K_{\chi ^\kappa _\rho }$ , $R(f[\omega ],\Xi ^2_{\tau ^\kappa _\rho }(f))$ .

Proof. Consider the game $H_R$ on $\kappa $ defined as follows:

Player 1 and Player 2 alternate playing ordinals from $\kappa $ . Player 1 plays $\alpha _{2n}$ and Player 2 plays $\beta _{2n + 1}$ as in the picture above for each $n \in \omega $ . Practically, one should regard Player 2 as playing a pair $\alpha _{2n + 1} \in \kappa $ and $x_n \in \omega $ such that $\pi ^{\kappa ,2}(\alpha _{2n + 1},x_n) = \beta _{2n + 1}$ . Let $g = \langle \alpha _0, \beta _1,\alpha _2,\beta _3, \ldots \rangle $ . Let $f = \langle \alpha _n : n \in \omega \rangle $ and $x = \langle x_n : n \in \omega \rangle $ . Player 2 wins if and only if the conjunction of the following holds:

  • For all $n \in \omega $ , $x_n \in \omega $ .

  • $R(\{\varphi _0(\mathfrak {G}_n(f)) : n \in \omega \},x)$ .

This game is determined by $\mathsf {AD}^+$ .

The claim is that Player 2 has a winning strategy in $H_R$ . For the sake of contradiction, suppose $\rho $ is a winning strategy for Player 1 in $H_R$ . Let $\sigma \in \mathscr {P}_{\omega _1}(\kappa )$ have the following two properties:

  1. (1) $\sigma $ is honest relative to the reliability witness $\vec \varphi $ .

  2. (2) $\rho (\emptyset ) \in \sigma $ . For all $k \in \omega $ , $\gamma _0,\ldots ,\gamma _{2k + 1} \in \sigma $ , $n_0, \ldots , n_{k} \in \omega $ ,

    $$ \begin{align*}\rho(\langle \gamma_0, \pi^{\kappa,2}(\gamma_1,n_0),\gamma_2, \pi^{\kappa,2}(\gamma_3,n_1), \ldots, \pi^{\kappa,2}(\gamma_{2k + 1},n_{k})\rangle) \in \sigma.\end{align*} $$

Let $x \in {{}^\omega \omega }$ be such that $R(\sigma ,x)$ . Let $h : \omega \rightarrow \sigma $ be a surjection onto $\sigma $ . Let $\tilde h : \omega \rightarrow \kappa $ be defined by $\tilde h(n) = \pi ^{\kappa ,2}(h(n),x(n))$ . Consider the run of $H_R$ where Player 1 uses $\rho $ and Player 2 uses $\rho ^2_{\tilde h}$ . Let $g = \rho * \rho ^2_{\tilde h}$ . Let $f(2n) = g(2n)$ and $f(2n + 1) = \pi ^{\kappa ,2}_0(g(2n + 1)) = h(n)$ . By (2), for all $n \in \omega $ , $f(2n) \in \sigma $ . Since for all $n \in \omega $ , $f(2n + 1) = h(n)$ and $h : \omega \rightarrow \sigma $ is a surjection, $f[\omega ] = \sigma $ . By (1), $f[\omega ]$ is honest. By the properties of the generic coding function $\mathfrak {G}$ (Fact 4.8), $\varphi _0(\mathfrak {G}_n(f)) = f(n)$ . Thus $\sigma = \{\varphi _0(\mathfrak {G}_n(f)) : n \in \omega \}$ . Note that $x(n) = \pi ^{\kappa ,2}_1(g(2n + 1))$ and $R(\sigma ,x)$ . This shows that Player 2 has won this run of $H_R$ , which contradicts $\rho $ being a winning strategy for Player 1.

Thus by the determinacy of $H_R$ , Player 2 has a winning strategy $\bar \rho $ . By the first condition for Player 2 winning, condition (1) must hold for $\bar \rho $ . Now suppose $h \in \mathscr {P}_{\omega _1}(\kappa )$ is such that $h[\omega ] \in K_{\chi ^\kappa _\rho }$ . Consider the run of $H_R$ where Player 1 plays by $\rho ^1_h$ and Player 2 plays by $\bar \rho $ . Let $g = \rho ^1_h * \bar \rho $ . Let $f : \omega \rightarrow \kappa $ be defined by $f(2n) = g(2n)$ and $f(2n + 1) = \pi ^{\kappa ,2}_0(g(2n + 1))$ . By the hypothesis that $h[\omega ] \in K_{\chi ^\kappa _{\bar \rho }}$ , $f(2n + 1) = \pi ^{\kappa ,0}(g(2n + 1)) \in h[\omega ]$ . Thus $f[\omega ] = \{f(n) : n \in \omega \} = h[\omega ],$ which is an honest set by the hypothesis that $h[\omega ] \in K_{\chi ^\kappa _{\bar \rho }}$ . By the properties of the generic coding function, $\varphi _0(\mathfrak {G}_n(f)) = f(n)$ . Thus $h[\omega ] = \{\varphi _0(\mathfrak {G}_n(f)) : n \in \omega \}$ . Let $x \in {{}^\omega \omega }$ be defined by $x(n) = \pi ^{\kappa ,2}_1(g(2n + 1))$ . Since $\bar \rho $ is a Player 2 winning strategy, $R(\{\varphi _0(\mathfrak {G}_n(f)) : n \in \omega \},x)$ holds or equivalently $R(h[\omega ], x)$ . Since $x = \Xi ^2_{\tau ^\kappa _{\bar \rho }}(h)$ , one has that $R(h[\omega ],\Xi ^2_{\tau ^\kappa _{\bar \rho }}(h))$ . This completes the proof.

In the following, one will focus on the supercompact measure on $\mathscr {P}_{\omega _1}(\omega _\omega )$ . One will develop first a coding of strategies on $\omega _\omega $ . The following objects will be fixed for the rest of the discussion concerning $\omega _\omega $ .

Definition 4.24. Fix a $\boldsymbol {\Pi }^1_2$ set W and a $\boldsymbol {\Delta }^1_3$ scale $\vec \varphi $ on W of length $\omega _\omega $ , which witnesses the reliability of $\omega _\omega $ . (This can be obtained by applying the scale property for $\boldsymbol {\Pi }^1_3$ on some complete $\boldsymbol {\Pi }^1_2$ set. More explicitly, one can let $W = \{x^\sharp : x \in \mathbb {R}\}$ and let $\vec \varphi $ be a modification of the sharp scale so that $\varphi _0 : W \rightarrow \omega _\omega $ is a surjection.) Let $\prec _n$ denote the prewellordering on W induced by $\varphi _n : W \rightarrow \omega _\omega $ . Note that $\prec _n \in \boldsymbol {\Delta }^1_3$ for all $n \in \omega $ . Fix a bijection $\pi ^{\omega _\omega ,<\omega } : \omega _\omega \rightarrow {}^{<\omega }(\omega _\omega )$ . Fix $U \subseteq \mathbb {R} \times \mathbb {R} \times \mathbb {R,}$ which is universal for $\boldsymbol {\Sigma }^1_3$ subsets of $\mathbb {R}^2$ .

Let $\mathsf {scode}$ be the set of $x \in \mathbb {R}$ so that the following holds:

  1. (1) For all $s \in {}^{<\omega }\omega _\omega $ , there exist $y,v \in \mathbb {R}$ such that $y \in W$ , $\pi ^{\omega _\omega ,<\omega }(\varphi _0(y)) = s$ , and $U(x,y,v)$ .

  2. (2) For all $y,z \in W$ , for all $v,w \in \mathbb {R}$ , if $\varphi _0(y) = \varphi _0(z)$ , $U(x,y,v)$ , and $U(x,z,w)$ , then $v,w \in W$ and $\varphi _0(v) = \varphi _0(w)$ .

For any $x \in \mathsf {scode}$ , $s \in {}^{<\omega }(\omega _\omega )$ , and $\alpha \in \omega _\omega $ , let $\rho _x(s) = \alpha $ if and only if there is a $y \in W$ and $v \in W$ so that $\pi ^{\omega _\omega ,<\omega }(\varphi _0(y)) = s$ , $\varphi _0(v) = \alpha $ , and $U(x,y,v)$ . By the two properties of $x \in \mathsf {scode}$ , $\rho _x$ is a well-defined function from ${}^{<\omega }(\omega _\omega )$ into $\omega _\omega $ (that is, $\rho _x$ is a strategy on $\omega _\omega $ ).

Let $\mathsf {scode}^*$ be the set of $x \in \mathbb {R}$ so that the following holds:

  1. (a) $x \in \mathsf {scode}$ .

  2. (b) For all $s \in {}^{<\omega }(\omega _\omega )$ so that $|s|$ is odd, for all $v \in \mathbb {R}$ , if $U(x,y,v)$ , then $\pi ^{\omega _\omega ,2}_1(\varphi _0(v)) \in \omega $ .

Note that if $x \in \mathsf {scode}^*$ , then $\Xi ^2_{\tau ^\kappa _\rho } : {}^\omega \kappa \rightarrow {{}^\omega \omega }$ .

Fact 4.25. For all strategies $\rho : {}^{<\omega }(\omega _\omega ) \rightarrow \omega _\omega $ , there is an $x \in \mathsf {scode}$ so that $\rho = \rho _x$ .

Proof. Define $R \subseteq W \times W$ by $R(y,v)$ if and only if $\rho (\pi ^{\omega _\omega ,<\omega }(\varphi _0(y))) = \varphi _0(v)$ . Applying the Moschovakis coding lemma to the pointclass $\boldsymbol {\Sigma }^1_3$ with the prewellordering $\varphi _0$ , there is an $S \subseteq R$ with $S \in \boldsymbol {\Sigma }^1_3$ so that for all $\beta \in \omega _\omega $ , there exists a $y \in W$ with $\varphi _0(y) = \beta $ and $v \in \mathbb {R}$ so that $S(y,v)$ . Since $\pi ^{\omega _\omega ,<\omega } : \omega _\omega \rightarrow {}^{<\omega }(\omega _\omega )$ is a bijection, this can be expressed also as: for all $s \in {}^{<\omega }(\omega _\omega )$ , there exist $y \in W$ and $v \in \mathbb {R}$ so that $\pi ^{\omega _\omega ,<\omega }(\varphi _0(y)) = s$ , $S(y,v)$ . Since $U \subseteq \mathbb {R} \times \mathbb {R} \times \mathbb {R}$ is universal for $\boldsymbol {\Sigma }^1_3$ subsets of $\mathbb {R}^2$ , there is some $x \in \mathbb {R}$ so that $U_x = S$ . By the previous observation and the fact that $U_x = S \subseteq R$ , one has properties (1) and (2) of Definition 4.24 and that $\rho _x = \rho $ .

One will need to make several complexity computations in order to use the Kunen–Martin theorem to bound the ultrapower $j_{\nu _{\omega _\omega }}$ . The closure of $\boldsymbol {\Delta }^1_4$ , $\boldsymbol {\Sigma }^1_4$ , and $\boldsymbol {\Pi }^1_4$ under $\omega _\omega $ -length unions will be helpful in making several complexity computations. This result, due to Harrington and Kechris, has analogs for other scaled pointclasses. For the results here, one can make even better complexity calculations using the Kechris–Martin theorem [Reference Kechris and Martin15, Corollary 1.6] to show $\boldsymbol {\Sigma }^1_3$ and $\boldsymbol {\Pi }^1_3$ are closed under $\omega _\omega $ -length unions and intersections. Jackson has extended the Kechris–Martin theorem throughout the projective hierarchy using the description theory [Reference Jackson14, Section 4.4]. However, these arguments are not known to generalize much further.

Fact 4.26. (Harrington–Kechris [Reference Harrington and Kechris11, Corollary 2.2] Assume $\mathsf {AD}$ . For all $n \in \omega $ , for all $\kappa < \boldsymbol {\delta }^1_{n}$ , $\boldsymbol {\Pi }^1_{n + 1}$ , $\boldsymbol {\Sigma }^1_{n + 1}$ , and $\boldsymbol {\Delta }^1_{n + 1}$ are closed under $\kappa $ -length union. In particular, $\boldsymbol {\Pi }^1_4$ , $\boldsymbol {\Sigma }^1_4$ , and $\boldsymbol {\Delta }^1_4$ are closed under $\omega _\omega $ -length unions.

Proof. The last statement follows from the first using $n = 3$ and the fact that $\boldsymbol {\delta }^1_3 = \omega _{\omega + 1}$ .

Fact 4.27. (Kechris and Moschovakis [Reference Kechris and Moschovakis16, Theorem 8.4] Assume $\mathsf {AD}$ . For all $n \in \omega $ , $\boldsymbol {\Delta }^1_{2n + 1}$ is closed under $\kappa $ -length unions and intersections for all $\kappa < \boldsymbol {\delta }^1_{2n + 1}$ . In particular, $\boldsymbol {\Delta }^1_3$ is closed under $\omega _\omega $ -length unions and intersections.

Fact 4.28. Assume $\mathsf {AD}$ . $\mathsf {scode}$ and $\mathsf {scode}^*$ are $\boldsymbol {\Delta }^1_4$ .

Proof. For each $s \in {}^{<\omega }(\omega _\omega )$ , let $A_s$ be the set $x \in \mathbb {R}$ so that there exist $y,v \in \mathbb {R}$ so that $y \in W$ , $\varphi _0(y) = (\pi ^{\omega _\omega ,<\omega })^{-1}(s)$ , and $U(x,y,v)$ . Note that $A_s$ is $\boldsymbol {\Sigma }^1_3$ since W is $\boldsymbol {\Pi }^1_2$ , $\varphi _0$ is a $\Delta ^1_3$ -norm, and U is $\boldsymbol {\Sigma }^1_3$ . In particular, $A_s$ is $\boldsymbol {\Delta }^1_4$ . Let $A = \bigcap \{A_s : s \in {}^{<\omega }(\omega _\omega )\},$ which is $\boldsymbol {\Delta }^1_4$ since $\boldsymbol {\Delta }^1_4$ is closed under $\omega _\omega $ -length intersection by Fact 4.26. (A is actually $\boldsymbol {\Sigma }^1_3$ since $\boldsymbol {\Sigma }^1_3$ is closed under $\omega _\omega $ -length intersections by the Kechris–Martin theorem.) Note that A is the set of $x \in \mathbb {R,}$ which satisfies Definition 4.24 property (1). Let B be the set of $x,$ which satisfies Definition 4.24 property (2). Since $W \in \boldsymbol {\Pi }^1_2$ , $U \in \boldsymbol {\Sigma }^1_3$ , and $\varphi _0$ is a $\Delta ^1_3$ norm, one has that B is $\boldsymbol {\Pi }^1_3$ . Since $\mathsf {scode} = A \cap B$ , $\mathsf {scode} \in \boldsymbol {\Delta }^1_4$ .

Let $X = \{\alpha \in \omega _\omega : \pi ^{\omega _\omega ,2}_1(\alpha ) \in \omega \}$ . For each $\alpha \in X$ and $s \in {}^{<\omega }(\omega _\omega )$ with $|s|$ odd, let $C_{\alpha ,s}$ be the set of x so that for all $y,v \in \mathbb {R}$ , if $v \in W$ , $\varphi _0(y) = (\pi ^{\omega _\omega ,<\omega })^{-1}(s)$ , and $U(x,y,v)$ , then $\varphi _0(v) = \alpha $ . Note that $C_{\alpha ,s}$ is $\boldsymbol {\Pi }^1_3$ . Let $C = \bigcap \{\bigcup \{C_{\alpha ,s} : \alpha \in X\} : s \in {}^{<\omega }(\omega _\omega ) \wedge |s| \text { is odd}\}$ . Since $\boldsymbol {\Delta }^1_4$ is closed under $\omega _\omega $ -length intersections and unions, $C \in \boldsymbol {\Delta }^1_4$ . Since $\mathsf {scode}^* = \mathsf {scode} \cap C$ , $\mathsf {scode}^*$ is $\boldsymbol {\Delta }^1_4$ .

Lemma 4.29. Assume $\mathsf {AD}$ .

  1. (1) Let $\mathsf {String} \subseteq \omega \times \mathbb {R} \times \mathbb {R}$ be defined by $\mathsf {String}(n,r,y)$ if and only if $y \in W$ , for all $m < n$ , $r^{[m]} \in W$ , and $\pi ^{\omega _\omega ,<\omega }(\varphi _0(y)) = \langle \varphi _0(r^{[0]}), \ldots , \varphi _0(r^{[n - 1]})\rangle $ (that is, $\pi ^{\omega _\omega ,<\omega }(\varphi _0(y))$ is the length n-string $\langle \varphi _0(r^{[0]}), \ldots , \varphi _0(r^{[n - 1]})\rangle $ ). $\mathsf {String}$ is $\boldsymbol {\Delta }^1_3$ .

  2. (2) Let $\mathsf {IntPart} \subseteq \mathbb {R} \times \omega $ be defined by $\mathsf {IntPart}(v,n)$ if and only if $v \in W$ and $\pi ^{\omega _\omega ,2}_1(\varphi _0(v)) = n$ . $\mathsf {IntPart} \in \boldsymbol {\Delta }^1_3$ .

  3. (3) Let $\mathsf {ONPart} \subseteq \mathbb {R} \times \mathbb {R}$ be defined by $\mathsf {ONPart}(v,w)$ if and only if $v \in W$ and $\pi ^{\omega _\omega ,2}_0(\varphi _0(v)) = \varphi _0(w)$ . $\mathsf {ONPart} \in \boldsymbol {\Delta }^1_3$ .

  4. (4) There is a $\Delta ^1_3$ relation $\mathsf {NormCompare} \subseteq \omega \times \omega \times \mathbb {R} \times \mathbb {R}$ so that for all $m,n \in \omega $ and $v,w \in \mathbb {R}$ , $\mathsf {NormCompare}(m,n,v,w)$ if and only if $v,w \in W$ and $\varphi _m(v) = \varphi _n(w)$ (where $\vec \varphi = \langle \varphi _n : n \in \omega \rangle $ come from the fixed reliability witness).

  5. (5) There is a $\boldsymbol {\Sigma }^1_3$ set $\mathsf {Honest} \subseteq \mathbb {R}$ so that $\mathsf {Honest}(r)$ if and only if for all $n \in \omega $ , $r^{[n]} \in W$ and $\{\varphi _0(r^{[n]}) : n \in \omega \}$ is honest relative to the reliability witness $\vec \varphi $ .

  6. (6) There is a $\boldsymbol {\Sigma }^1_3$ relation $\mathsf {Run}_{\boldsymbol {\Sigma }^1_3} \subseteq \mathbb {R} \times \mathbb {R}$ and a $\boldsymbol {\Pi }^1_3$ relation $\mathsf {Run}_{\boldsymbol {\Pi }^1_3}$ so that if $x \in \mathsf {scode}$ , then $\mathsf {Run}_{\boldsymbol {\Sigma }^1_3}(x,r)$ if and only if $\mathsf {Run}_{\boldsymbol {\Pi }^1_3}(x,r)$ if and only if $\langle \varphi _0(r^{[n]}) : n \in \omega \rangle $ is a run according to $\rho _x$ used as a strategy for Player 2.

  7. (7) There is a $\boldsymbol {\Sigma }^1_3$ relation $\mathsf {Closed}_{\boldsymbol {\Sigma }^1_3} \subseteq \mathbb {R} \times \mathbb {R}$ and $\boldsymbol {\Pi }^1_3$ relation $\mathsf {Closed}_{\boldsymbol {\Pi }^1_3} \subseteq \mathbb {R} \times \mathbb {R}$ with the property that whenever $x \in \mathsf {scode}$ , $\mathsf {Closed}_{\boldsymbol {\Sigma }^1_3}(x,r)$ if and only if $\mathsf {Closed}_{\boldsymbol {\Pi }^1_3}(x,r)$ if and only if for all $n \in \omega $ , $r^{[n]} \in W$ and for all for all $s \in {}^{<\omega }(\{\varphi _0(r^{[n]}) : n \in \omega \})$ , $\rho _x(s) \in \{\varphi _0(r^{[n]}) : n \in \omega \}$ .

  8. (8) There is a $\boldsymbol {\Sigma }^1_3$ relation $\mathsf {fClosed}_{\boldsymbol {\Sigma }^1_3} \subseteq \mathbb {R} \times \mathbb {R}$ and $\boldsymbol {\Pi }^1_3$ relation $\mathsf {fClosure}_{\boldsymbol {\Pi }^1_3} \subseteq \mathbb {R} \times \mathbb {R}$ with the property that whenever $x \in \mathsf {scode}$ , $\mathsf {fClosed}_{\boldsymbol {\Sigma }^1_3}(x,r)$ if and only if $\mathsf {fClosed}_{\boldsymbol {\Pi }^1_3}(x,r)$ if and only if for all $n \in \omega $ , $r^{[n]} \in W$ and for all $s \in {}^{<\omega }(\{\varphi _0(r^{[n]}) : n \in \omega \})$ , $\chi _{\rho _x}^{\omega _\omega }(s) \in \{\varphi _0(r^{[n]}) : n \in \omega \}$ .

Proof.

  1. (1) For each $s \in {}^{<\omega }(\omega _\omega )$ , let $A_s$ be the set of $(|s|,r,y)$ such that $y \in W$ , $\varphi _0(y) = (\pi ^{\omega _\omega ,<\omega })^{-1}(s)$ , and for all $m < n$ , $r^{[m]} \in W$ and $\varphi _0(r^{[m]}) = s(m)$ . Note that $A_s \in \boldsymbol {\Delta }^1_3$ and $\mathsf {String} = \bigcup \{A_s : s \in {}^{<\omega }(\omega _\omega )\}$ . $\mathsf {String} \in \boldsymbol {\Delta }^1_3$ since $\boldsymbol {\Delta }^1_3$ is closed under $\omega _\omega $ -length unions by Fact 4.27.

  2. (2) For each $\alpha \in \omega _\omega $ and $n \in \omega $ , let $V_{\alpha ,n} = \{(v,n) : v \in W \wedge \varphi _0(v) = (\pi ^{\omega _\omega ,2})^{-1}((\alpha ,n))\}$ . Since $\varphi _0$ is a $\boldsymbol {\Delta }^1_3$ -norm, $V_{\alpha ,n} \in \boldsymbol {\Delta }^1_3$ . Then $\mathsf {IntPart} = \bigcup \{V_{\alpha ,n} : \alpha \in \omega _\omega \wedge n \in \omega \},$ which is $\boldsymbol {\Delta }^1_3$ since $\boldsymbol {\Delta }^1_3$ is closed under $\omega _\omega $ -length unions.

  3. (3) For each $\alpha ,\beta < \omega _\omega $ , let $(v,w) \in A_{\alpha ,\beta }$ if and only if $\varphi _0(v) = \pi ^{\omega _\omega ,2}(\alpha ,\beta )$ and $\beta = \varphi _0(w)$ . $A_{\alpha ,\beta }$ is $\Delta ^1_3$ . $\mathsf {ONPart} = \bigcup \{A_{\alpha ,\beta } : \alpha ,\beta < \omega _\omega \},$ which is $\boldsymbol {\Delta }^1_3$ since $\boldsymbol {\Delta }^1_3$ is closed under $\omega _\omega $ -length unions.

  4. (4) Let $m,n \in \omega $ and $\alpha < \omega _\omega $ . If $\alpha $ is greater than or equal to the rank of either $\varphi _m$ or $\varphi _n$ , then let $A_{m,n,\alpha } = \emptyset $ . If $\alpha $ less than the rank of both $\varphi _m$ and $\varphi _n$ , then let $A_{m,n,\alpha } = \{(m,n,v,w) : \varphi _m(v) = \alpha \wedge \varphi _n(w) = \alpha \}$ . $A_{m,n,\alpha }$ is $\boldsymbol {\Delta }^1_3$ since all the norms in $\vec \varphi $ are $\boldsymbol {\Delta }^1_3$ norms. Then $\mathsf {NormCompare} = \bigcup \{A_{m,n,\alpha } : m,n \in \omega \wedge \alpha < \omega _\omega \},$ which is $\boldsymbol {\Delta }^1_3$ since $\boldsymbol {\Delta }^1_3$ is closed under $\omega _\omega $ -length unions.

  5. (5) Note that $r \in \mathsf {Honest}$ if and only if for all $n \in \omega $ , there exists $w \in W$ so that $\varphi _0(w) = \varphi _0(r^{[n]})$ and for all $k \in \omega $ , there exists $j \in \omega $ such that $\mathsf {NormCompare}(0,k,r^{[j]},w)$ . Since $\mathsf {NormCompare}$ is $\boldsymbol {\Delta }^1_3$ , $\mathsf {Honest}$ is $\boldsymbol {\Sigma }^1_3$ .

  6. (6) Let $\mathsf {Run}_{\boldsymbol {\Sigma }^1_3}(x,r)$ if and only if for all $n \in \omega $ , $r^{[n]} \in W$ and there exist $y, v \in \mathbb {R}$ so that $\mathsf {String}(2n + 1, r, y)$ , $U(x,y,v)$ , and $\varphi _0(v) = \varphi _0(r^{[2n + 1]})$ . $\mathsf {Run}_{\boldsymbol {\Sigma }^1_3}$ is $\boldsymbol {\Sigma }^1_3$ and if $x \in \mathsf {scode}$ , then $\mathsf {Run}_{\boldsymbol {\Sigma }^1_3}(x,r)$ has the intended meaning stated above.

    Let $\mathsf {Run}_{\boldsymbol {\Pi }^1_3}(x,r)$ if and only if for all $n \in \omega $ , $r^{[n]} \in W$ and for all $y,v \in \mathbb {R}$ , if $\mathsf {String}(2n + 1, r, y)$ and $U(x,y,v)$ , then $\varphi _0(v) = \varphi _0(r^{[2n + 1]})$ . $\mathsf {Run}_{\boldsymbol {\Pi }^1_3}$ is $\boldsymbol {\Pi }^1_3$ and if $x \in \mathsf {scode}$ , then $\mathsf {Run}_{\boldsymbol {\Pi }^1_3}(x,r)$ has the intended meaning.

  7. (7) This is a similar and simpler than the argument shown next for (8).

  8. (8) Define $\mathsf {fClosed}_{\boldsymbol {\Pi }^1_3}(x,r)$ if and only if the conjunction of the following holds:

    • For all $n \in \omega $ , $r^{[n]} \in W$ .

    • For all $n \in \omega $ , for all $t,y,v,v_0 \in \mathbb {R}$ , if the conjunction of the following holds:

      • For all $k < n$ , there exists $i \in \omega $ , $\varphi _0(t^{[k]}) = \varphi _0(r^{[i]}).$

      • $\mathsf {String}(n,t,y)$ .

      • $U(x,y,v).$

      • $\mathsf {ONPart}(v,v_0)$ .

      then there exists a $j \in \omega $ , $\varphi _0(v_0) = \varphi _0(r^{[j]})$ .

    Note that $\mathsf {fClosed}_{\boldsymbol {\Pi }^1_3} \in \boldsymbol {\Pi }^1_3$ .Define $\mathsf {fClosed}_{\boldsymbol {\Sigma }^1_3}(x,r)$ if and only if the conjunction of the following holds:

    • For all $n \in \omega $ , $r^{[n]} \in W$ .

    • For all $n \in \omega $ and function $\ell : n \rightarrow \omega $ , there exist $j \in \omega $ and $t,y,v,v_0 \in \mathbb {R}$ so that the conjunction of the following holds:

      • For all $k < n$ , $t^{[k]} = r^{[\ell (k)]}$ .

      • $\mathsf {String}(n,t,y)$ .

      • $U(x,y,v).$

      • $\mathsf {ONPart}(v,v_0)$ .

      • $\varphi _0(v_0) = \varphi _0(r^{[j]})$ .

    Note that $\mathrm {fClosed}_{\boldsymbol {\Sigma }^1_3}$ is $\boldsymbol {\Sigma }^1_3$ .If $x \in \mathsf {scode}$ , then $\mathsf {fClosed}_{\boldsymbol {\Sigma }^1_3}$ and $\mathsf {fClosed}_{\boldsymbol {\Pi }^1_3}$ have the intended meanings.

Fact 4.30. Assume $\mathsf {AD}$ . Suppose $x \in \mathsf {scode}^*$ . Let A be the set of $f \in {}^\omega (\omega _\omega )$ so that $f[\omega ] \in K_{\chi ^{\omega _\omega }_{\rho _x}}$ . Then $\Xi ^2_{\tau ^{\omega _\omega }_{\rho _x}}[A]$ is $\boldsymbol {\Sigma }^1_3$ (note that since $x \in \mathsf {scode}^*$ , $\Xi ^2_{\tau ^{\omega _\omega }_{\rho _x}}[A]$ is a set of reals).

Proof. Observe that $u \in \Xi ^2_{\tau ^{\omega _\omega }_{\rho _x}}[A]$ if and only if there exist $r,t \in \mathbb {R}$ so that the conjunction of the following holds:

  • $\mathsf {fClosed}_{\boldsymbol {\Sigma }^1_3}(x,r).$

  • $\mathsf {Honest}(r)$ .

  • For all $n \in \omega $ , $t^{[2n]} = r^{[n]}$ .

  • $\mathsf {Run}_{\boldsymbol {\Sigma }^1_3}(x,t)$ .

  • For all $n \in \omega $ , $\mathsf {IntPart}(t^{[2n + 1]},u(n))$ .

The above expression is $\boldsymbol {\Sigma }^1_3$ and it works because $x \in \mathsf {scode}^*$ (and note that $\mathsf {scode}^* \subseteq \mathsf {scode}$ ).

Fact 4.31. (Steel [Reference Steel24], [Reference Jackson14, Theorem 2.28]) Assume $\mathsf {AD}$ and $\mathsf {DC}_{\mathbb {R}}$ . If $\kappa < \Theta $ is a limit ordinal, then there is a surjective norm $\psi : P \rightarrow \kappa ,$ which is $\delta $ -Suslin bounded for all $\delta < {\mathrm {cof}}(\kappa )$ , which means that for all $A \subseteq P$ that are $\delta $ -Suslin, $\sup (\varphi [A]) < \kappa $ .

Fact 4.32. Assume $\mathsf {AD}^+$ . Let $\kappa < \Theta $ with ${\mathrm {cof}}(\kappa )> \omega _\omega $ . Let $\Phi : \mathscr {P}_{\omega _1}(\omega _\omega ) \rightarrow \kappa $ . Then there is an $A \in \nu _{\omega _\omega }$ so that $\sup (\Phi [A]) < \kappa $ .

Proof. Fix $\kappa < \Theta $ with ${\mathrm {cof}}(\kappa )> \omega _\omega $ . By Fact 4.31, let $\psi : P \rightarrow \kappa $ be a surjective $\omega _\omega $ -Suslin bounded prewellordering. Fix $\Phi : \mathscr {P}_{\omega _1}(\omega _\omega ) \rightarrow \kappa $ . Let $R \subseteq \mathscr {P}_{\omega _1}(\omega _\omega ) \times \mathbb {R}$ be defined by $R(\sigma ,p)$ if and only if $\Phi (\sigma ) = \psi (p)$ . Applying Fact 4.23, there is a strategy $\rho $ so that the following holds:

  1. (1) For all odd length $s \in {}^{<\omega }(\omega _\omega )$ , $\tau ^{\omega _\omega }_{\rho }(s) \in \omega $ .

  2. (2) For all $f \in {}^\omega (\omega _\omega )$ so that $f[\omega ] \in K_{\chi ^{\omega _\omega }_\rho }$ , $R(f[\omega ], \Xi ^2_{\tau ^{\omega _\omega }_\rho }(f))$ .

By Fact 4.25, there is an $x \in \mathsf {scode}$ so that $\rho _x = \rho $ . Moreover, $x \in \mathsf {scode}^*$ by condition (1) above. Let B be the set of $f \in {}^\omega (\omega _\omega )$ so that $f[\omega ] \in K_{\chi ^{\omega _\omega }_{\rho _x}}$ . By condition (2), for any $f \in B$ , $R(f[\omega ],\Xi ^2_{\tau ^{\omega _\omega }_{\rho _x}}(f))$ and thus $\Xi ^2_{\tau ^{\omega _\omega }_{\rho _x}}(f) \in P$ by the definition of R. Thus $\Xi ^2_{\tau ^{\omega _\omega }_{\rho _x}}[B] \subseteq P$ and $\Xi ^2_{\tau ^{\omega _\omega }_{\rho _x}}[B]$ is $\boldsymbol {\Sigma }^1_3$ (and hence $\omega _\omega $ -Suslin) by Fact 4.30. Since $\psi $ is a $\omega _\omega $ -Suslin bounded norm, there is a $\delta < \kappa $ so that $\psi [\Xi ^2_{\rho _x,1}[B]] \subseteq \delta $ . $K_{\chi ^{\omega _\omega }_{\rho _x}}\in \nu _{\omega _\omega }$ by Fact 4.15. Let $\sigma \in K_{\chi ^{\omega _\omega }_{\rho _x}}$ . Let $f : \omega \rightarrow \sigma $ be any surjection and thus $f[\omega ] = \sigma $ . Note that $f \in B$ . Therefore by (2), $R(\sigma ,\Xi ^2_{\tau ^{\omega _\omega }_{\rho _x}}(f))$ . This means $\Phi (\sigma ) = \psi (\Xi ^2_{\tau ^{\omega _\omega }_{\rho _x}}(f))$ . Since $\psi (\Xi ^2_{\tau ^{\omega _\omega }_{\rho _x}}(f)) \in \Xi ^2_{\tau ^{\omega _\omega }_{\rho _x}}[B]$ , one has that $\psi (\Xi ^2_{\tau ^{\omega _\omega }_{\rho _x}}(f)) < \delta $ . So $\Phi (\sigma ) < \delta $ . This shows that $\sup (\Phi [K_{\chi ^{\omega _\omega }_{\rho _x}}]) \leq \delta < \kappa $ .

Definition 4.33. Let $\mathsf {scode}^+$ consists of those $x \in \mathbb {R}$ so that the following hold:

  1. (1) $x \in \mathsf {scode}^*$ .

  2. (2) For all $f \in {}^\omega (\omega _\omega )$ so that $f[\omega ] \in K_{\chi ^{\omega _\omega }_{\rho _x}}$ , $\Xi ^2_{\tau ^{\omega _\omega }_{\rho _x}}(f) \in W$ (where recall W is the underlying set of norms that form the reliability witness $\vec \varphi $ ).

  3. (3) For all $f_0,f_1 \in {}^\omega (\omega _\omega )$ so that $f_0[\omega ], f_1[\omega ] \in K_{\chi ^{\omega _\omega }_{\rho _x}}$ and $f_0[\omega ] = f_1[\omega ]$ , then $\varphi _0(\Xi ^2_{\tau ^{\omega _\omega }_{\rho _x}}(f_0)) = \varphi _0(\Xi ^2_{\tau ^{\omega _\omega }_{\rho _x}}(f_1))$ .

If $x \in \mathsf {scode}^+$ , then let $\Phi _x : K_{\chi ^{\omega _\omega }_{\rho _x}} \rightarrow \omega _\omega $ be defined by $\Phi _x(\sigma ) = \varphi _0(\Xi ^2_{\tau ^{\omega _\omega }_{\rho _x}}(f))$ for any $f : \omega \rightarrow \sigma ,$ which is a surjection. The conditions of the definition of $\mathsf {scode}^+$ imply that $\Phi _x$ is a well-defined function independent of the choice of f which surjects onto $\sigma $ .

Fact 4.34. Assume $\mathsf {AD}^+$ . For any $\Phi : \mathscr {P}_{\omega _1}(\omega _\omega ) \rightarrow \omega _\omega $ , there is an $x \in \mathsf {scode}^+$ so that $[\Phi ]_{\nu _{\omega _\omega }} = [\Phi _x]_{\nu _{\omega _\omega }}$ .

Proof. This was shown in the proof of Fact 4.32. (Replace the $\psi : P \rightarrow \kappa $ of the proof of Fact 4.32 with $\varphi _0 : W \rightarrow \omega _\omega $ .) (Moreover, if one inspects the payoff set for Player 2 in the game $H_R$ for the relevant relation R from Fact 4.32, one can even strengthen Definition 4.33 condition (2) to say that for all $f \in {}^\omega (\omega _\omega )$ , $\Xi ^2_{\tau ^{\omega _\omega }_{\rho _x}}(f) \in W$ .)

Fact 4.35. Assume $\mathsf {AD}$ . $\mathsf {scode}^+$ is $\boldsymbol {\Delta }^1_4$ .

Proof. Note that $x \in \mathsf {scode}^+$ if and only if the conjunction of the following hold:

  • $x \in \mathsf {scode}^*$ .

  • For all $r,t,u \in \mathbb {R}$ , if the conjunction of the following hold:

    • $\mathsf {Honest}(r)$ .

    • $\mathsf {fClosed}_{\boldsymbol {\Sigma }^1_3}(x,r)$ .

    • For all $n \in \omega $ , $t^{[2n]} = r^{[n]}$ .

    • For all $n \in \omega $ , $\mathsf {IntPart}(t^{[2n + 1]}, u(n)).$

    • $\mathsf {Run}_{\boldsymbol {\Sigma }^1_3}(x,t)$ ,

    then $u \in W$ .

  • For all $r_0,t_0,u_0, r_1, t_1, u_1 \in \mathbb {R}$ , if the conjunction of the following hold:

    • $\mathsf {Honest}(r_0)$ and $\mathsf {Honest}(r_1)$ .

    • $\mathsf {fClosed}_{\boldsymbol {\Sigma }^1_3}(x,r_0)$ . $\mathsf {fClosed}_{\boldsymbol {\Sigma }^1_3}(x,r_1)$ .

    • For all $n \in \omega $ , $(t_0)^{[2n]} = (r_0)^{[n]}$ and $(t_1)^{[2n]} = (r_1)^{[n]}$ .

    • For all $n \in \omega $ , $\mathsf {IntPart}((t_0)^{[2n + 1]}, u_0(n))$ and $\mathsf {IntPart}((t_0)^{[2n + 1]}, u_0(n))$ .

    • $\mathsf {Run}_{\boldsymbol {\Sigma }^1_3}(x,t_0)$ and $\mathsf {Run}_{\boldsymbol {\Sigma }^1_3}(x,t_1)$ ,

    • For all $m \in \omega $ , there exists $n \in \omega $ so that $\varphi _0((r_0)^{[m]}) = \varphi _0((r_1)^{[n]})$ . For all $m \in \omega $ , there exists $n \in \omega $ so that $\varphi _0((r_1)^{[m]}) = \varphi _0((r_0)^{[n]})$ .

    then $\varphi _0(u_0) = \varphi _0(u_1)$ .

The first point is $\boldsymbol {\Delta }^1_4$ since $\mathsf {scode}^* \in \boldsymbol {\Delta }^1_4$ . The second and third points are $\boldsymbol {\Pi }^1_3$ . The entire expression is $\boldsymbol {\Delta }^1_4$ .

Fact 4.36. (Kunen–Martin Theorem) Assume $\mathsf {AC}^{\mathbb {R}}_\omega $ . Every $\kappa $ -Suslin wellfounded relation on $\mathbb {R}$ has length less than $\kappa ^+$ .

Fact 4.37. (Becker [Reference Becker1, Theorem 4.2]) Assume $\mathsf {AD}^+$ . Let $\alpha < \boldsymbol {\delta }^1_3 = \omega _{\omega + 1}$ and $\nu _\alpha $ be the unique supercompact measure on $\mathscr {P}_{\omega _1}(\alpha )$ . Then $j_{\nu _\alpha }(\boldsymbol {\delta }^1_4) = j_{\nu _\alpha }(\omega _{\omega + 2}) = \boldsymbol {\delta }^1_4 = \omega _{\omega + 2}$ .

Proof. Note that these ultrapowers are wellfounded by Fact 4.22. For all ${\alpha < \boldsymbol {\delta }^1_3 = \omega _{\omega + 1}}$ , $\nu _\alpha $ is Rudin–Keisler reducible to $\nu _{\omega _\omega }$ by Fact 4.20 and therefore $j_{\nu _\alpha }(\boldsymbol {\delta }^1_4) \leq j_{\nu _{\omega _\omega }}(\boldsymbol {\delta }^1_4)$ . Thus it suffices to show that $j_{\nu _{\omega _\omega }}(\boldsymbol {\delta }^1_4) = \boldsymbol {\delta }^1_4$ .

The representatives of ordinals below $j_{\nu _{\omega _\omega }}(\boldsymbol {\delta }^1_4)$ are functions of the form ${\Phi : \mathscr {P}_{\omega _1}(\omega _\omega ) \rightarrow \boldsymbol {\delta }^1_4}$ . Since $\boldsymbol {\delta }^1_4$ is regular, Fact 4.32 implies that $\Phi $ is $\nu _{\omega _\omega }$ -almost equal to a function which is strictly bounded below $\boldsymbol {\delta }^1_4$ . Thus $j_{\nu _{\omega _\omega }}(\boldsymbol {\delta }^1_4) = \sup \{j_{\nu _{\omega _\omega }}(\beta ) : \beta < \boldsymbol {\delta }^1_4\}$ . To prove the theorem, it suffices to show that $j_{\nu _{\omega _\omega }}(\beta ) < \boldsymbol {\delta }^1_4$ for all $\beta < \boldsymbol {\delta }^1_4$ .

Let $\beta < \boldsymbol {\delta }^1_4 = \omega _{\omega + 2}$ . Since $\boldsymbol {\delta }^1_3 = \omega _{\omega + 1}$ , let $\psi _\beta : \boldsymbol {\delta }^1_3 \rightarrow \beta $ be a surjection. For each $\Phi : \mathscr {P}_{\omega _1}(\omega _\omega ) \rightarrow \boldsymbol {\delta }^1_3$ , let $\tilde \Phi : \mathscr {P}_{\omega _1}(\omega _\omega ) \rightarrow \beta $ be defined by $\tilde \Phi (\sigma ) = \psi (\Phi (\sigma ))$ . For every $\Upsilon : \mathscr {P}_{\omega _1}(\omega _\omega ) \rightarrow \beta $ , there is a $\Phi : \mathscr {P}_{\omega _1}(\omega _\omega ) \rightarrow \boldsymbol {\delta }^1_3$ so that $\tilde \Phi = \Upsilon $ . Thus $\Psi : j_{\nu _{\omega _\omega }}(\boldsymbol {\delta }^1_3) \rightarrow j_{\nu _{\omega _\omega }}(\beta )$ defined by $\Psi ([\Phi ]_{\nu _{\omega _\omega }}) = [\tilde \Phi ]_{\nu _{\omega _\omega }}$ for any $\Phi : \mathscr {P}_{\omega _1}(\omega _\omega ) \rightarrow \boldsymbol {\delta }^1_3$ is a well-defined surjection. Since $\boldsymbol {\delta }^1_4$ is a cardinal, it suffices to show that $j_{\nu _{\omega _\omega }}(\boldsymbol {\delta }^1_3) < \boldsymbol {\delta }^1_4$ .

Since $\boldsymbol {\delta }^1_3$ is regular, Fact 4.32 again implies $j_{\nu _{\omega _\omega }}(\boldsymbol {\delta }^1_3) = \sup \{j_{\nu _{\omega _\omega }}(\gamma ) : \gamma < \boldsymbol {\delta }^1_3\}$ . Since $\boldsymbol {\delta }^1_4$ is regular, it suffices to show that $j_{\nu _{\omega _\omega }}(\gamma ) < \boldsymbol {\delta }^1_4$ for all $\gamma < \boldsymbol {\delta }^1_3$ . Since $\boldsymbol {\delta }^1_3 = \omega _{\omega + 1}$ , the same argument from the previous paragraph shows that $j_{\nu _{\omega _{\omega }}}(\omega _\omega )$ surjects onto $j_{\nu _{\omega _\omega }}(\gamma )$ for all $\gamma < \boldsymbol {\delta }^1_3$ . Finally, it has been shown that to prove the theorem it suffices to show $j_{\nu _{\omega _\omega }}(\omega _\omega ) < \boldsymbol {\delta }^1_4$ .

Define a relation $\mathsf {compare} \subseteq \mathbb {R} \times \mathbb {R}$ as follows: $\mathsf {compare}(x,y)$ if and only there exists a $z \in \mathbb {R}$ such that the conjunction of the following hold:

  1. (1) $x,y \in \mathsf {scode}^+$ and $z \in \mathsf {scode}$ .

  2. (2) For all $r, t_0, t_1, u_0, u_1 \in \mathbb {R}$ , if the conjunction of the following hold:

    • $\mathsf {Honest}(r)$ .

    • $\mathsf {Closed}_{\boldsymbol {\Sigma }^1_3}(z,r)$ , $\mathsf {fClosed}_{\boldsymbol {\Sigma }^1_3}(x,r)$ , and $\mathsf {fClosed}_{\boldsymbol {\Sigma }^1_3}(y,r)$ .

    • For all $n \in \omega $ , $(t_0)^{[2n]} = (t_1)^{[2n]} = r^{[n]}$ .

    • For all $n \in \omega $ , $\mathsf {IntPart}((t_0)^{[2n + 1]}, u_0(n))$ and $\mathsf {IntPart}((t_1)^{[2n + 1]}, u_1(n))$ .

    • $\mathsf {Run}_{\boldsymbol {\Sigma }^1_3}(x,t_0)$ and $\mathsf {Run}_{\boldsymbol {\Sigma }^1_3}(y,t_1)$ .

    then $\varphi _0(u_0) < \varphi _0(u_1)$ .

Observe that (1) is $\boldsymbol {\Delta }^1_4$ and (2) is $\boldsymbol {\Pi }^1_3$ . Thus $\mathsf {compare}$ is $\boldsymbol {\Sigma }^1_4$ .

Claim 1: $\mathsf {compare}(x,y)$ if and only if $x,y \in \mathsf {scode}^+$ and $[\Phi _x]_{\nu _{\omega _\omega }} < [\Phi _y]_{\nu _{\omega _\omega }}$ .

To see Claim: $(\Rightarrow )$ Let z witness the existential quantifier in $\mathsf {compare}(x,y)$ . Note $K_{\chi ^{\omega _\omega }_{\rho _x}} \cap K_{\chi ^{\omega _\omega }_{\rho _y}} \cap K_{\rho _z} \in \nu _{\omega _\omega }$ . Let $\sigma \in K_{\chi ^{\omega _\omega }_{\rho _x}} \cap K_{\chi ^{\omega _\omega }_{\rho _y}} \cap K_{\rho _z}$ . By definition, this means that $\sigma $ is honest and closed under $\chi ^{\omega _\omega }_{\rho _x}$ , $\chi ^{\omega _\omega }_{\rho _x}$ , and $\rho _z$ . Let $f : \omega \rightarrow \sigma $ be any surjection. Let $g_x = \rho ^1_f * \rho _x$ and $g_y = \rho ^1_f * \rho _y$ . Let $r, t_0, t_1$ be such that for all $n \in \omega $ , $\varphi _0(r^{[n]}) = f(n)$ , $r^{[n]} = (t_0)^{[2n]}$ , $r^{[n]} = (t_1)^{[2n]}$ , $\varphi _0((t_0)^{[n]}) = g_x(n)$ , and $\varphi _0((t_1)^{[n]}) = g_y(n)$ . For all $n \in \omega $ , let $u_0(n) = \pi ^{\omega _\omega ,2}_1(\varphi _0((t_0)^{[2n + 1]}))$ and $u_1(n) = \pi ^{\omega _\omega ,2}_1(\varphi _0((t_1)^{[2n + 1]}))$ . r, $t_0$ , $t_1$ , $u_0$ , $u_1$ satisfy the hypothesis of the conditional in statement (2). Thus ${\varphi _0(u_0) < \varphi _0(u_1)}$ . Since $u_0 = \Xi ^2_{\tau ^{\omega _\omega }_{\rho _x}}(f)$ and $u_1 = \Xi ^2_{\tau ^{\omega _\omega }_{\rho _y}}(f)$ , one has that $\Phi _x(\sigma ) = \varphi _0(u_0) < \varphi _0(u_1) = \Phi _y(\sigma )$ by definition. Since $\sigma \in K_{\chi ^{\omega _\omega }_{\rho _x}} \cap K_{\chi ^{\omega _\omega }_{\rho _y}} \cap K_{\rho _z} \in \nu _{\omega _\omega }$ was arbitrary, this shows that $[\Phi _x]_{\nu _{\omega _\omega }} < [\Phi _y]_{\nu _{\omega _\omega }}$ .

$(\Leftarrow )$ Suppose $[\Phi _x]_{\nu _{\omega _\omega }} < [\Phi _y]_{\nu _{\omega _\omega }}$ . The set $A = \{\sigma \in \mathscr {P}_{\omega _1}(\omega _\omega ) : \Phi _x(\sigma ) < \Phi _y(\sigma )\} \in \nu _{\omega _\omega }$ . By Fact 4.16, there is a strategy $\rho $ so that $K_\rho \subseteq A$ . By Fact 4.25, there is a $z \in \mathsf {scode}$ so that $\rho _z = \rho $ . By much of the same argument as before, z witnesses the existential to show that $\mathsf {compare}(x,y)$ holds. This establishes the claim.

Define an equivalence relation $\sim $ on $\mathsf {scode}^+$ by $x \sim y$ if and only if $[\Phi _x]_{\nu _{\omega _\omega }} = [\Phi _y]_{\nu _{\omega _\omega }}$ . Let be the set of equivalence classes of $\sim $ . For $X,Y \in H$ , define $X < Y$ if and only if for any $x \in X$ and $y \in Y$ , $[\Phi _x]_{\nu _{\omega _\omega }} < [\Phi _y]_{\nu _{\omega _\omega }}$ . Observe that $(H,<)$ order embeds into $j_{\nu _{\omega _\omega }}(\omega _\omega )$ by the well-defined map $\Lambda (X) = [\Phi _x]_{\nu _{\omega _\omega }}$ for any $x \in X$ . This shows that $(H,<)$ is a wellordering. Hence by using the claim, $\mathsf {compare}$ is a wellfounded relation whose length corresponds to the ordertype of $(H,<)$ . By Fact 4.34, every $\Phi : \mathscr {P}_{\omega _1}(\omega _\omega ) \rightarrow (\omega _\omega )$ has an $x \in \mathsf {scode}^+$ so that $[\Phi ]_{\nu _{\omega _\omega }} = [\Phi _x]_{\nu _{\omega _\omega }}$ . This shows that the ordertype of $(H,<)$ is exactly $j_{\nu _{\omega _\omega }}(\omega _\omega )$ . Hence the length of $\mathsf {compare}$ is exactly $j_{\nu _{\omega _\omega }}(\omega _\omega )$ . Since $\mathsf {compare}$ is a wellfounded $\boldsymbol {\Sigma }^1_4$ and hence $\boldsymbol {\delta }^1_3 = \omega _{\omega + 1}$ Suslin relation, the Kunen–Martin theorem states that the length of $\mathsf {compare}$ is less than $(\boldsymbol {\delta }^1_3)^+ = (\omega _{\omega + 1})^+ = \omega _{\omega + 2} = \boldsymbol {\delta }^1_4$ . Thus $j_{\nu _{\omega _\omega }}(\omega _\omega ) < \boldsymbol {\delta }^1_4$ . This completes the proof.

Theorem 4.38. Assume $\mathsf {AD}^+$ . Let $\langle A_\alpha : \alpha < \boldsymbol {\delta }^1_3\rangle $ be such that $\bigcup _{\alpha < \boldsymbol {\delta }^1_3} A_\alpha = {\mathscr {P}(\boldsymbol {\delta }^1_4)}$ . Then there is an $\alpha < \boldsymbol {\delta }^1_3$ so that $\neg (|A_\alpha | \leq |{}^{<\boldsymbol {\delta }^1_4}\boldsymbol {\delta }^1_4|)$ .

Proof. Suppose ${\mathscr {P}(\boldsymbol {\delta }^1_4)} = \bigcup _{\alpha < \boldsymbol {\delta }^1_3} A_\alpha $ and $|A_\alpha | \leq |{}^{<\boldsymbol {\delta }^1_4}\boldsymbol {\delta }^1_4|$ for all $\alpha < \boldsymbol {\delta }^1_3$ . $\boldsymbol {\delta }^1_3$ is a Suslin cardinal and hence reliable. By Fact 4.37, the hypothesis of Theorem 4.11 holds. Thus $|{\mathscr {P}(\boldsymbol {\delta }^1_4)}| = |\bigcup _{\alpha < \boldsymbol {\delta }^1_3} A_\alpha | \leq |{}^{<\boldsymbol {\delta }^1_4}\boldsymbol {\delta }^1_4|$ . $\boldsymbol {\delta }^1_4$ is a weak partition cardinal and hence a measurable cardinal. Thus $\boldsymbol {\delta }^1_4$ does not inject into ${\mathscr {P}(\gamma )}$ for any $\gamma < \boldsymbol {\delta }^1_4$ . So $|{}^{<\boldsymbol {\delta }^1_4}\boldsymbol {\delta }^1_4| < |{\mathscr {P}(\boldsymbol {\delta }^1_4)}|$ by Fact 2.9. This is a contradiction.

This argument can be generalized to the suitable analog at higher projective ordinals.

Theorem 4.39. Assume $\mathsf {AD}^+$ . Let $n \in \omega $ . Let $\langle A_\alpha : \alpha < \boldsymbol {\delta }^1_{2n + 1}\rangle $ be such that $\bigcup _{\alpha < \boldsymbol {\delta }^1_{2n + 1}} A_\alpha = {\mathscr {P}(\boldsymbol {\delta }^1_{2n + 2})}$ . Then there is an $\alpha < \boldsymbol {\delta }^1_{2n + 1}$ so that $\neg (|A_\alpha | \leq |{}^{<\boldsymbol {\delta }^1_{2n + 2}}\boldsymbol {\delta }^1_{2n + 2}|)$ .

Funding

The first author was support by NSF grants DMS-1945592 and DMS-1800323 and FWF grants I6087 and Y1498. The second author was supported by NSF grant DMS-1800323. The third author was supported by NSF grants DMS-1855757 and DMS-1945592.

References

REFERENCES

Becker, H., $\mathsf{AD}$ and the supercompactness of ${\aleph}_1$ . The Journal of Symbolic Logic, vol. 46 (1981), no. 4, pp. 822842.Google Scholar
Becker, H. and Jackson, S., Supercompactness within the projective hierarchy . The Journal of Symbolic Logic, vol. 66 (2001), no. 2, pp. 658672.Google Scholar
Caicedo, A. E. and Ketchersid, R., A trichotomy theorem in natural models of ${\mathsf{AD}}^{+}$ , Set Theory and its Applications (L. Babinkostova, A. E. Caicedo, S. Geschke, and M. Scheepers, editors), Contemporary Mathematics, 533, American Mathematical Society, Providence, 2011, pp. 227258.Google Scholar
Chan, W., Cardinality of the set of bounded subsets of a cardinal. In preparation.Google Scholar
Chan, W., Definable combinatorics of strong partition cardinals. In preparation.Google Scholar
Chan, W., An introduction to combinatorics of determinacy , Trends in Set Theory (S. Coskey and G. Sargsyan, editors), Contemporary Mathematics, 752, American Mathematical Society, Providence, 2020, pp. 2175. MR 4132099Google Scholar
Chan, W., Jackson, S., and Trang, N., Boldface $\mathsf{GCH}$ below the first uncountable limit cardinal. In preparation.Google Scholar
Chan, W., Jackson, S., and Trang, N., A search for the cofinality of the power set of the first uncountable cardinal. In preparation.Google Scholar
Chan, W., Jackson, S., and Trang, N., The size of the class of countable sequences of ordinals . Transactions of the American Mathematical Society, vol. 375 (2022), no. 3, pp. 17251743.Google Scholar
Chan, W., Jackson, S., and Trang, N., Almost everywhere behavior of functions according to partition measures . Forum of Mathematics Sigma, vol. 12 (2024), Article no. e16.Google Scholar
Harrington, L. A. and Kechris, A. S., On the determinacy of games on ordinals . Annals of Mathematical Logic, vol. 20 (1981), no. 2, pp. 109154.Google Scholar
Hjorth, G., A dichotomy for the definable universe . The Journal of Symbolic Logic, vol. 60 (1995), no. 4, pp. 11991207.Google Scholar
Jackson, S., The weak square property . The Journal of Symbolic Logic, vol. 66 (2001), no. 2, pp. 640657.Google Scholar
Jackson, S., Structural consequences of $\mathsf{AD}$ , Handbook of Set Theory (A. Kanamori and M. Foreman, editors), vols. 1, 2, 3, Springer, Dordrecht, 2010, pp. 17531876. MR 2768700Google Scholar
Kechris, A. S. and Martin, D. A., On the theory of ${\varPi}_3^1$ sets of reals, II, Ordinal Definability and Recursion Theory: The Cabal Seminar (A. S. Kechris, B. Lowe, and J. R. Steel, editors), vol. 3, Lecture Notes in Logic, 43, Association for Symbolic Logic, Ithaca, 2016, pp. 200219. MR 3469170Google Scholar
Kechris, A. S. and Moschovakis, Y. N., Notes on the Theory of Scales Cabal Seminar 76–77 (Caltech-UCLA Logic Seminar, 1976–77), Lecture Notes in Mathematics, 689, Springer, Berlin, 1978, pp. 153.Google Scholar
Kechris, A. S. and Woodin, W. H., Generic codes for uncountable ordinals, partition properties, and elementary embeddings , Games, Scales, and Suslin Cardinals. The Cabal Seminar (A. S. Kechris, B. Lowe, and J. R. Steel, editors), vol. 1, Lecture Notes in Logic, 31, Association for Symbolic Logic, Chicago, 2008, pp. 379397. MR 2463619Google Scholar
Ketchersid, R., More structural consequences of $\mathsf{AD}$ , Set theory and its Applications (L. Babinkostova, A. E. Caicedo, S. Geschke, and M. Scheepers, editors), Contemporary Mathematics, 533, American Mathematical Society, Providence, 2011, pp. 71105. MR 2777745Google Scholar
Kleinberg, E. M., Infinitary Combinatorics and the Axiom of Determinateness, Lecture Notes in Mathematics, 612, Springer-Verlag, Berlin, 1977.Google Scholar
Larson, P. B., Extensions of the Axiom of Determinacy, University Lecture Series, 78, American Mathematical Society, Providence, 2023.Google Scholar
Larson, P. B. and Zapletal, J., Geometric Set Theory, Mathematical Surveys and Monographs, 248, American Mathematical Society, Providence, 2020.Google Scholar
Moschovakis, Y. N., Descriptive Set Theory, second ed., Mathematical Surveys and Monographs, 155, American Mathematical Society, Providence, 2009.Google Scholar
Solovay, R. M., The independence of $\mathsf{DC}$ from $\mathsf{AD}$ , Large Cardinals, Determinacy and Other Topics. The Cabal Seminar (A. S. Kechris, B. Lowe, and J. R. Steel, editors), vol. 4, Lecture Notes in Logic, 49, Cambridge University Press, Cambridge, 2021, pp. 6695. MR 4567353Google Scholar
Steel, J. R., Closure Properties of Pointclasses, Cabal Seminar 77–79 (Proceedings. Caltech-UCLA Logic Seminars 1977–79), Lecture Notes in Mathematics, 839, Springer, Berlin, 1981, pp. 147163.Google Scholar
Steel, J. R., ${HOD}^{L\left(\mathbb{R}\right)}$ is a core model below $\varTheta$ . The Bulletin of Symbolic Logic, vol. 1 (1995), no. 1, pp. 7584.Google Scholar
Steel, J. R. and Woodin, W. H., HOD as a core model , Ordinal Definability and Recursion Theory: The Cabal Seminar (A. S. Kechris, B. Lowe, and J. R. Steel, editors), vol. 3, Lecture Notes in Logic, 43, Association for Symbolic Logic, Ithaca, 2016, pp. 257345. MR 3469173Google Scholar
Steel, J. and Trang, N., ${\mathsf{AD}}^{+}$ , derived models, and ${\varSigma}_1$ -reflection. https://math.berkeley.edu/~steel/papers/AD+reflection.pdf Google Scholar
Woodin, W. H., $\mathsf{AD}$ and the uniqueness of the supercompact measures on $\mathcal{P}_{\omega_1}(\lambda)$ , Large Cardinals, Determinacy and Other Topics. The Cabal Seminar (A. S. Kechris, B. Lowe, and J. R. Steel, editors), vol. 4, Lecture Notes in Logic, 49, Cambridge University Press, Cambridge, 2021, pp. 136140. MR 4567357Google Scholar